diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzbluy" "b/data_all_eng_slimpj/shuffled/split2/finalzzbluy" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzbluy" @@ -0,0 +1,5 @@ +{"text":"\\section*{Supplemental Material}\n\nThis supplemental material conveys additional results on the weak coupling regime, the effects of a harmonic trap, correlation functions and mutual information in the strong coupling phase separated regime, and a derivation of the spinful $PXP$ Hamiltonian at half and full filling.\n\n\n\n\\section{Weak coupling Rydberg dressed regime}\n\nTo obtain the weak-coupling Hamiltonian, see Eq. \\eqref{eqn:cos} and above it, we linearize the tight-binding dispersion around $\\pm k_F$. Without interactions, the bosonized Hamiltonian takes the form:\n\\begin{equation}\n H^{nonint}_\\eta = \\frac{1}{2\\pi} \\int dx \\left[ v (\\pi \\Pi_\\eta (x))^2 + v (\\nabla \\phi_\\eta(x))^2 \\right],\n\\end{equation}\nwhere $v_F= \\frac{d \\varepsilon}{dk}_{k=k_F}=\\frac{2t}{a}\\sin(k_F a)$, where $a$ is the lattice constant of the order of $\\alpha$. Next we transform the lattice interaction Eq.~\\eqref{eq:rydbergint} to the momentum space:\n\\begin{equation}\n V_d(q) = \\frac{1}{N}\\sum_n \n e^{-i q a n } V_d(a n).\n\\end{equation}\nOne can write $V_d(a n) = \\int \\frac{d k}{2\\pi} e^{i k a n} V_C(k) = \\frac{1}{L}\\sum_{k=\\frac{2\\pi l}{L}} e^{i k a n} V_C(k)|_{L\\to\\infty}$, where $V_C(k) = \\int dx V_d(x) e^{ikx}$ is defined in the continuum (C) limit $a\\rightarrow 0$. Using the Poisson summation formula $\\sum_n e^{i n g} = N \\sum_m \\delta_{g-2\\pi m}$ one obtains:\n\\begin{equation}\n V_d(q) = \\frac{1}{L}\\sum_m V_C\\left(q+\\frac{2\\pi m}{a}\\right).\n\\end{equation}\nAs $V_C(k)$ decays as $e^{-|k|r_c\/2}$ for $r_c\\gg a$ the $m\\neq0$ terms can be neglected and $V_d(q)\\approx \\frac{1}{L} V_c(q)$ used in the main text is obtained.\n\n\n\n\n\n\nApart from the spin-spin correlations shown in Fig.~\\ref{fig:weak}(a), we also calculate the density-density correlation \n$\\langle n_{L\/2 \\pm x} n_{L\/2} \\rangle$\nfor the two (positive and negative backscattering) phases in Fig.~\\ref{fig:rhorho}. \nAs the charge sector is always gapless the density-density correlation decays as power-law in both cases.\nThe density-density and spin-spin power-law exponents in the repulsive backscattering region are both $-(K_\\rho + K_\\sigma)$. \nWe extracted the exponent for $N_p = 120$ from the spin-spin correlator (dashed line in Fig.~\\ref{fig:weak}(a)) and plotted a dashed line with the same slope in Fig.~\\ref{fig:rhorho}. \nOne can confirm that the power (slope) of the two matches well in the large $x$ limit.\n\n\\begin{figure}[b!]\n \\centering\n \\includegraphics[width=0.9\\linewidth]{rhocorr4.png}\n \\caption{The density-density correlation function for $N_p = 70$ (blue) and $N_p = 120$ (yellow) particles in a $L = 200$ chain.\n All parameters are identical to that of Fig.~\\ref{fig:weak}.\n The density-density correlation follows a power-law scaling on both cases. \n }\n \\label{fig:rhorho}\n\\end{figure}\n\n\\section{Effect of trap potential}\n\n\nThe calculations in the main text focus on an optical lattice but ignores the presence of the harmonic trap potential.\nIn real experiments, the harmonic trap is inevitable and will affect the previously shown results. While we expect this to have a weak effect on the weak coupling phases, we need to confirm that the phase separated regime in Fig.~\\ref{fig:strong}(b) survives. We include a harmonic trap by adding to the Hamiltonian a term $\\sum_{r,\\sigma} V_r c_{r\\sigma}^{\\dag}c_{r\\sigma}$ with $V_r = \\frac{1}{2} m\\omega^2 r^2$.\n\n\\begin{figure}[t!]\n \\centering\n \\includegraphics[width=0.49\\linewidth]{trapN1.png}\n \\includegraphics[width=0.49\\linewidth]{trapEE1.png}\n \\includegraphics[width=0.49\\linewidth]{trapN2.png}\n \\includegraphics[width=0.49\\linewidth]{trapEE2.png}\n \\caption{\n Left accumulated particle $\\mathcal{N}_L(x)$ as a function of position and the bipartite entanglement entropy in a $L = 400$ chain in the presence of a harmonic trap potential.\n (a) $N_p = 167$ and $\\frac{1}{2} m\\omega^2 x_{\\textrm max}^2 = t$, (b) $N_p = 163$ and $\\frac{1}{2} m\\omega^2 x_{\\textrm max}^2 = 2t$.\n The yellow dashed lines are the linear fit near the center ($x = 200$) and the deviation from this line indicates different filling (and thus different phase) away from the center.\n }\n \\label{fig:trap}\n\\end{figure}\n\nHere, we take the experimental setup in Ref.~\\onlinecite{Guardado-Sanchez2021} with $t = h \\times 1.7$ kHz and $a = 752$ nm. \nFig.~\\ref{fig:trap} is the case of $N_p = 167, 163$ particles in a $L=400$ chain with a trap potential $\\frac{1}{2} m\\omega^2 x_{\\textrm max}^2 = t, 2t$, respectively with $x_{\\textrm max}=L\/2$.\nThis corresponds to a trap frequency of order $10^2$Hz, which is safely larger than the typical harmonic trap. \n\nIt is clear that the slope (average particle density) at the center is different from that of the two boundaries. \nThe bipartite entanglement entropy also shows distinct behavior between center and boundaries, and the entanglement is low at the center regions, following the trend seen in Fig.~\\ref{fig:strong}(c). Here, the center region is the magnetic domain wall and the charge like puddles are to its side. \nFrom these results, while the overall structure of the phase separated state is modified by the trap \nwe conclude that the phase separated regime of the model will not be prohibited by the harmonic trap in experiments and we expect can be directly measured. \n\n\n\\section{Correlations in the phase separated regime}\n\n\n\\begin{figure*}[t!]\n \\centering\n \\includegraphics[width=0.32\\linewidth]{PS_rhorho.png}\n \\includegraphics[width=0.325\\linewidth]{PS_sigsig.png}\n \\includegraphics[width=0.30\\linewidth]{MI.png}\n \\caption{\n Correlation functions for the same parameters in Fig.~\\ref{fig:strong}(b,c). \n ($L=400$, $N_p=170$ in the strong coupling regime) \n (a) The density-density correlation function where $x_0 = 87$ [blue], $212$ [yellow], $324$ [green] which are the centers of the $c=1$ puddles in Fig.~\\ref{fig:strong}(c).\n (b) The spin-spin correlation function where $x_0 = 154$ [blue], $263$ [yellow] which are in between the $c=1$ puddles.\n (c) The mutual information (MI) between two points $x$, $y$ in the system.\n The MI shows three puddles which corresponds to the three $c=1$ regions, and the information spread is blocked by the magnetized region. \n }\n \\label{fig:MI}\n\\end{figure*}\n\nIn this section, we further investigate the phase separated state in the context of correlations to provide additional results that imply the presence of Luttinger liquid puddles.\nIn the main text Fig.~\\ref{fig:strong}(c), we claimed that the three regions with $\\bullet$$\\bullet$$\\circ$$\\circ$$\\circ$ patterns have a central charge of $c=1$ which results from the gapless charge degrees of freedom and is inherited from the weak coupling limit of the problem. \nWe check this directly calculating the density-density correlation function and observing its functional form. \n\nIn Fig.~\\ref{fig:MI}(a) we plot the density-density correlation function for the same parameters in Fig.~\\ref{fig:strong}(b,c), where the center $x_0$ is chosen as the center of the three $c=1$ regions. \nOne can observe slow power-law decay of the correlation, although the exponent is hard to extract due to the small size of each region.\nIn contrast, the spin-spin correlation function decays exponentially, even in the magnetized region. \nIn Fig.~\\ref{fig:MI}(b) shows the fast decay of such correlation, where $x_0$ is the center of the magnetized region (in between the $c=1$ regions).\n\nFinally, we plot the mutual information (MI) between two points $x$, $y$ in the system. \nThe MI between two subsets $A$ and $B$ of the system is defined as~\\cite{Amico08}:\n\\begin{align}\n \\mathcal{I}(A, B) = \\mathcal{S}(A) + \\mathcal{S}(B) - \\mathcal{S}(A\\cup B),\n\\end{align}\nwhere $\\mathcal{S}(A)$ is the entanglement entropy (or von Neumann entropy) introduced in the main text.\nFig.~\\ref{fig:MI}(c) plots the MI between all pair of points in the system, and shows that the MI is large only between points within the same $c=1$ region. \nThis shows the natural result that mutual information is large in gapless regions, and also the interesting feature that the gapped regions serve as a barrier of information spread between the gapless $c=1$ puddles.\n\n\\section{Bosonization analysis at strong coupling}\n\n\nWe now present details of the bosonization discussion regarding phases for arbitrary commensurate filling. For a filling equal to $\\frac{p}{q}$, $p,q$ being mutually prime integers, an additional (umklapp) term appears in the low-energy theory, corresponding to simultaneous scattering of $q$ fermions from $-k_F$ to $k_F$. This can be written as \\cite{schulz_mott} \n\\begin{equation}\n H_{um}=U_{um}\\int dx \\cos(q\\sqrt{2} \\varphi_\\rho) \\cos^q(\\sqrt{2}\\varphi_\\sigma)\n\\end{equation}, where for weak coupling $U_{um}\\sim U (U\/t)^{q-2}$ \\cite{schulz_mott,Giamarchi}. For even $q$, the most relevant part of $\\cos^q(\\sqrt{2}\\varphi_\\sigma)$ is just a constant, such that $U_{um}$ is relevant for $K_\\rho<4\/q^2$, leading to a charge gap in that case. Apart from half filling, this implies that the density degree of freedom can remain gapless below a threshold value of $U$. For odd $q$, $\\cos(\\sqrt{2}\\varphi_\\sigma)$ has to be kept. At weak coupling, if $V_C(2k_F)<0$ this term is fixed $\\cos(\\sqrt{2}\\varphi_\\sigma)=\\pm1$ and the same condition for $K_\\rho$ is valid , while for $V_C(2k_F)>0$ one can use the fixed-point value $K_\\sigma=1$ \\cite{schulz_mott} such that $K_\\rho< 3\/q^2$ is required. In the latter case, a gap will open simultaneously in the density and spin excitations. The regions of $1\/5$ filling of Fig. \\ref{fig:strong} (b,c) actually conform to weak coupling expectations. In that case $k_F r_c\\approx 5$ such that $V_C(2k_F)\/V_0 \\approx - 0.16$ and a spin gap is expected, while the density gap requires strong interactions $K_\\rho<0.16$. Note that the observed strong modulation of density occurs at $2k_F$, same as for a weak-coupling density wave. The \n$\\odot$%\n$\\odot$%\n$\\circ$%\n$\\circ$%\n$\\circ$%\n$\\circ$ regions, on the other hand, has filling $1\/4$ with a less stringent requirement for the charge gap $K_\\rho<0.25$, allowing to understand the development of a charge gap there. While in weak coupling, a spin gap would have been also expected due to $V_C(2k_F)\/V_0\\approx-0.026$, its absence points to inapplicability of the continuum weak coupling analysis for $1\/4$ filling here, possibly due to extremely small value of $V_C(2k_F)\/V_0$ and finite size effects. The above arguments also suggest that for 1\/6 filling of Fig.~\\ref{fig:strong}(a), there will be a spin gap $V_C(2k_F)\/V_0\\approx-0.2$ and no charge gap ($K_\\rho<0.11$ is required), consistent with the numerical result. The charge modulation has the period $6$ (see Fig.~\\ref{fig:strong2}(a)), corresponding to the usual $2k_F=2\\pi\/6$ Luttinger liquid correlations \\cite{Giamarchi}, in contrast to the period $3$ ($4k_F$) modulation expected in the insulating state \\cite{schulz_mott,Giamarchi}.\nThe bipartite entanglement entropy (see Fig.~\\ref{fig:strong2}(b)) also shows clear $c=1$ scaling. \n\n\n\n\\begin{figure}[t!]\n \\centering\n \\includegraphics[width=0.49\\linewidth]{rhocorr5.png}\n \\includegraphics[width=0.49\\linewidth]{PS_EE3.png}\n \\caption{\n (a) The power-law scaling density-density correlation for the system in Fig.~\\ref{fig:strong}(a), $N_p = 140$. (b) The bipartite entanglement entropy as a function of the cut position. The solid line is the CFT scaling form (Eq.~\\eqref{eq:ee}) with $c=1$. \n }\n \\label{fig:strong2}\n\\end{figure}\n\n\\section{Optical tweezers and the $PXP$ limit}\n\nIn the limit of no tunneling (i.e. focusing on the optical tweezer set up as opposed to an optical lattice) and strong coupling we determine the effective many-body Hamiltonian, recovering the $PXP$ model in the limit of a fully filled Fermi sea and identify a magnetic $PXP$ model with a direct exchange interaction in the limit of half-filling.\n\n\nThe $PXP$ limit assumes a strong mixing of ground and Rydberg state $\\Omega \\gg \\Delta$. In this case, the dominant interaction is the repulsion between adjacent Rydberg atoms, that is also larger than $\\Omega$. Essentially, the interaction forbids atoms in Rydberg state to occupy neighboring sites. For spinless bosons, Rydberg Hamiltonian reduces then to the $PXP$ model, $H = \\sum_i P_{i-1}X_i P_{i+1}$~\\cite{sun2008,olmos2012,turner2018}. \n$P_i = (1-Z_i)\/2$ is a projector to the ground state. \nThe typical $PXP$ model studied intensively in the context of quantum many-body scars and slow dynamics focused on bosonic Rydberg atoms.\nHere, we consider fermionic Rydberg atoms which introduces the spin degrees of freedom to the original problem. \n\nFirst, consider a chain of Rydberg atoms with two particles (spin up and down) per site. \nThe $X_i$ operator connects the $|g\\uparrow, g\\downarrow\\rangle$ state with the $(|R\\uparrow,g\\downarrow\\rangle + |g\\uparrow,R\\downarrow\\rangle)\/\\sqrt{2}$ state and thus this model reduces to the original $PXP$ model.\nThis is reasonable in the sense that each site has two fermions which can be considered as a bosonic degree of freedom. \n\nNext, we consider a Rydberg atom chain with the filling of one particle per site. Neglecting the coupling of the laser to the spin degree of freedom (which is appropriate for $s$-like states), the Hamiltonian reduces to the original $PXP$ model with additional spin degeneracy, i.e.\n\\begin{align}\n H_{PXP}^{spin} = \\sum_i P_{i-1} X_i P_{i+1},\n\\end{align}\nwhere $X = | R \\uparrow\\rangle\\langle g\\uparrow| + | R \\downarrow\\rangle\\langle g\\downarrow| + \\textrm{h.c.}$ is the Pauli matrix in the $eg$-space, $P = | g \\uparrow\\rangle\\langle g\\uparrow| + | g \\downarrow\\rangle\\langle g\\downarrow|$ is the projection to the ground state.\n\n\n\nThe spin degeneracy, however, can be lifted by interactions due to the exchange process. \nBoth exchange and supexchange interactions depend on wavefunction overlap \\cite{auerbach2012interacting} and can be of the same order. Therefore, one can expect that: (1) they decrease with distance fast (2) they are larger for the Rydberg state.\nAs the Rydberg interaction energy prohibits nearest neighbor atoms to be both in the Rydberg state, the next possibility is to either have spin-spin interactions with a neighboring ground-state atom, or a next-neighbor Rydberg atom. Combining both contributions, the spin exchange term can be written as:\n\\begin{equation}\n\\begin{gathered}\n H_{exch}^{spin} = \\sum_i J_{Rg} \\vec{S}_i \\cdot \\vec{S}_{i+1} (P_i Q_{i+1} + Q_i P_{i+1}) \n \\\\\n + J_{RR} \\vec{S}_i \\cdot \\vec{S}_{i+2} Q_i Q_{i+2},\n \\end{gathered}\n\\end{equation}\nwhere $Q = 1 - P$.\n\n\nFor the direct exchange mechanism one can provide the following estimates for the couplings $J_{Rg}$ and $J_{RR}$. The Rydberg-Rydberg exchange coupling takes the form\\cite{auerbach2012interacting} :\n\\begin{equation}\n\\begin{gathered}\n J_{RR} = \\frac{1}{2} \\int d^3 {\\bf r} d^3{\\bf r}' V_R(|{\\bf r} - {\\bf r}'|) \\psi^*_R({\\bf r}) \\psi_R({\\bf r}')\\times \n \\\\\n \\times\\psi_R^*({\\bf r}'+2 a \\hat{x} \\bf) \\psi_R({\\bf r}+2 a \\hat{x} \\bf),\n \\end{gathered}\n\\end{equation}\nwhere $\\hat{x}$ is a unit vector along the chain direction and $V_R(|{\\bf r} - {\\bf r}'|) \\propto 1\/|{\\bf r} - {\\bf r}'|^6$, and $\\psi_{g,R}({\\bf r})$ correspond to ground and excited state atomic wavefunctions. One observes then that the integral is dominated by ${\\bf r} \\approx {\\bf r}'$ and is determined by the overlap between Rydberg wavefunctions displaced by $2a$. As the size of the Rydberg wavefunction scales with principal quantum number as $n^2$, the integral can be estimated to scale (taking the integrand value in the middle between two atoms) as $\\sim \\exp\\left( - 2 \\frac{2a}{a_B^* n^2}\\right)$, where $a_B^*$ is the Bohr radius for the atom. \n\nThe Rydberg-ground exchange coupling is determined by:\n\\begin{equation}\n\\begin{gathered}\n J_{Rg} = \\frac{1}{2} \\int d^3 {\\bf r} d^3{\\bf r}' V_{Rg}(|{\\bf r} - {\\bf r}'|) \\psi^*_R({\\bf r}) \\psi_R({\\bf r}')\\times\n \\\\\n \\times\\psi_g^*({\\bf r}'+a \\hat{x} \\bf) \\psi_g({\\bf r}+ a \\hat{x} \\bf),\n \\end{gathered}\n\\end{equation}\nwhere $V_{Rg}(|{\\bf r} - {\\bf r}'|)$ is the interaction between Rydberg and ground state atom, that is likely to be much weaker than the Rydberg interaction. The scaling of this integral with $n$ can be estimated analogously to be $\\sim \\exp\\left( - \\frac{a}{a_B^* n^2}- \\frac{a}{a_B^*}\\right)$. \nUnfortunately, in typical tweezer setups \\cite{bernien2017} $a\\gg a_B^* n^2$, suggesting that the exchange interactions are likely quite small for the current set ups.\n\n\n\n\\end{appendix}\n\n\\end{document}","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nGiven a pointed space $X$, denote by $\\mathcal E(X)$ the group of\n(based) self homotopy equivalences, i.e., the group of\nautomorphisms of $X$ in the pointed homotopy category. From now on\nwe shall consider connected complexes of finite type $X$ which are\neither finite or with finitely many non trivial homotopy groups.\nWe denote by $\\dim X=N$ its topological or homotopical dimension.\nUnless explicitly stated otherwise, all spaces will be of this\nkind.\n\nAlthough the computation of $\\mathcal E(X)$ is known to be a hard task,\nthere are two classical and key results that impose to this group\nimportant structural constraints:\n\nOn one hand, a theorem of Sullivan \\cite[Theorem 10.3]{su} and\nWilkerson \\cite[Theorem B]{wi} states that $\\mathcal E(X)$ is finitely\npresented. This was originally proved for simply connected spaces\nand later on generalized to virtually nilpotent spaces by Dror,\nDwyer and Kan \\cite[Theorem 1.1]{ddk}. The main step in the proof\nis to show that $\\mathcal E(X_\\mathbb Q)$ is an algebraic group and that\n$\\mathcal E(X)$ is commensurable with an arithmetic subgroup of\n$\\mathcal E(X_\\mathbb Q)$. As a consequence, it can be shown that there exists\na finite bound for the finite orders of elements of $\\mathcal E(X)$.\n\nOn the other hand we have the following theorem due to Dror and\nZabrodsky:\n\n\\begin{theorem}\\label{principal}{\\em \\cite[Theorem B]{d-z}} Let $G$ be a subgroup of\n$\\mathcal E(X)$ which acts nilpotently on $\\pi_{\\le N}(X)$. Then $G$ is\nitself nilpotent. In particular, $\\mathcal E_{ \\sharp }^m(X)$ is nilpotent.\n\\end{theorem}\n\nRecall that, for $0\\le m\\le\\infty$, $\\mathcal E_{ \\sharp }^m(X)$ is the\ndistinguished subgroup of $\\mathcal E(X)$ formed by those classes\ninducing the identity on the homotopy groups up to $m$. In other\nwords, $$ \\mathcal E_{ \\sharp }^m(X)=\\text{ker}\\bigl(\\mathcal E(X)\\longrightarrow\n\\Pi_{i\\le m}\\text{aut}\\,\\pi_iX\\bigr). $$ If $\\dim X=N$ we shall\ndenote $\\mathcal E_{ \\sharp }^N(X)$ simply by $\\mathcal E_{ \\sharp }(X)$.\n\n\n Here, we present a slightly different proof for this well known result\nin which we use a broader study of self homotopy equivalences in\nthe homotopy category ${\\cal L}^*$ of (based) spaces with local\ncoefficients. Recall (see \\cite[Chap.VI]{white} for instance) that\nobjects in this category are pairs $(X,{\\cal M})$ in which $X$ is a\n(based) topological space and ${\\cal M}=\\{M_x\\}_{x\\in X}$ is a local\ncoefficient system in $X$. On the other hand, a morphism\n$(f,\\Theta)\\colon (X,{\\cal M})\\to(Y,{\\cal H})$ is a pair formed by a\nbased map $f\\colon X\\to Y$ and a morphism $\\Theta\\colon\nf^*{\\cal H}\\to{\\cal M}$ of local coefficient. By $f^*{\\cal H}$ we denote,\nas usual, the local coefficient system on $X$ induced by $f$,\ni.e., $(f^*{\\cal H})_x=H_{f(x)}$. For each $x\\in X$ we shall denote\nby $\\Theta_x\\colon H_{f(x)}\\to M_x$ the corresponding group\nmorphism at $x$. After considering the appropriate homotopy\nnotion, one obtains the homotopy category ${\\cal L}^*$. The\ngroup of self homotopy equivalences of an object $(X,{\\cal M})\\in\n{\\cal L}^*$ shall be denoted by ${\\mathcal E}(X;\\calm)$. Then, we prove:\n\n\n\n\n\n\n\\begin{theorem}\\label{local} Let $X$ be a finite Postnikov piece\nand let $G\\subset {\\mathcal E}(X;\\calm)$ be a subgroup which acts nilpotently\non both $\\pi_*(X)$ and ${\\cal M}$. Then, $G$ acts nilpotently on $H^*(X;{\\cal M})$.\n\\end{theorem}\n\n\nAt the sight of Theorem \\ref{principal}, and taking into account\nthe bound of finite orders of elements of $\\mathcal E(X)$ plus the\nexistence of a ``fracture lemma\" for this group (see \\cite[Theorem\n8.2]{Rutter}), it has been of interest to study whether $\\mathcal E_{ \\sharp }(X)$\nsatisfies the same structural restrictions when taking\n$p$-localization, $p$-completion or considering $\\mathcal E_{ \\sharp p }(X)$. This\ndenotes the subgroup of $\\mathcal E(X)$ formed by those classes which\ninduce the identity on the homotopy groups of $X$ with\ncoefficients on $\\mathbb Z\/p$, up to the dimension of $X$, i.e., $$\n\\mathcal E_{ \\sharp p }(X)=\\text{ker}\\bigl(\\mathcal E(X)\\longrightarrow \\Pi_{\\le\nN}\\text{aut}\\,\\pi_i(X;\\mathbb Z\/p)\\bigr). $$ As examples of this, we\nmention two interesting results for a given nilpotent space $X$:\nMaruyama proved \\cite[Theorem 0.1]{ma} that\n$\\mathcal E_{ \\sharp }(X)_{(p)}=\\mathcal E_{ \\sharp }^N(X_{(p)})$ while, on the other hand, M{\\o}ller\nshowed \\cite[Theorem 4.3]{mo} that\n$$\\mathcal E_{ \\sharp }(X_{\\mathbb Z_p})=Ext(\\mathbb Z\/p^\\infty,\\mathcal E_{ \\sharp }(X))=\n\\mathcal E_{ \\sharp }(X_p{\\displaystyle\\hat{}}). $$ Here and henceforth,\n$(-)_{\\mathbb Z_p}$ denotes $H_*(-;\\mathbb Z\/p)$-localization while\n$(-)_{(p)}$ and $(-)_p{\\displaystyle\\hat{}}$ are the classical\nlocalization and completion on the prime $p$.\n\nIn this paper we plan to continue this investigation extending\nTheorem \\ref{principal} above, considering a subgroup of $\\mathcal E(X)$\nwhich acts nilpotently in the homotopy groups of $X$ localized,\ncompleted or with coefficients in $\\mathbb Z\/p$. Concerning this purpose\nwe prove:\n\n\n\n\\begin{theorem}\\label{dos} Assume that $\\pi_1(X)$ is a nilpotent group and let $G$ be a subgroup of $\\mathcal E(X)$ which acts\nnilpotently on $\\pi_{\\le N}(X)_{(p)}$, for $p$ any prime number\nand $0$. If the nilpotency orders of all these actions are bounded\nby a fixed integer, then G is nilpotent.\n\\end{theorem}\n\\begin{remark} {\\em Observe that in the theorem above the condition of $\\pi_1(X)$ being nilpotent is essential. Otherwise,\nchoose any finite simple group $G$ which is known to be\ngenerically trivial, i.e., $G_{(p)}=\\{1\\}$ for $p$ any prime\nnumber or zero. On the other hand observe that the map $G\\to \\text{aut}\\,\nG$ given by inner automorphisms is a monomorphism. Indeed, its\nkernel is the center of $G$ which is trivial since $G$ is simple.\nThis inclusion renders the non nilpotent group $G$ as a subgroup\nof $\\mathcal E\\bigl(K(G,1)\\bigr)$ which acts nilpotently in the\nlocalized homotopy group.\n\nIn a similar way, we may even produce an example of a solvable,\nnon nilpotent group, of homotopy equivalences acting nilpotently\nin the localized homotopy groups of the space. Consider the\nsymmetric group $\\Sigma_3$ and observe that\n${\\Sigma_3}_{(2)}=\\mathbb Z\/2$ while ${\\Sigma_3}_{(p)}=1$ for $p\\not=2$.\nAgain, $\\Sigma_3\\subset\\mathcal E\\bigl(K(\\Sigma_3,1)\\bigr)$ is a\nsolvable non nilpotent group acting nilpotently on any\n${\\Sigma_3}_{(p)}$.}\n\\end{remark}\n\nA more subtle and slightly different situation is given when\nconsidering nilpotent actions of subgroups of self homotopy\nequivalences on the Frattini factor of the homotopy groups. Recall\nthat given a group $G$, the Frattini subgroup $\\Phi(G)$ is the\nintersection of all maximal proper subgroups of $G$. The quotient\n$G\/\\Phi(G)$ is called the Frattini factor.\n\n\\begin{theorem}\\label{uno} Assume that\n$\\pi_{\\le N}(X)$ is a finite nilpotent group and let $G$ be a\nsubgroup of $\\mathcal E(X)$ which acts nilpotently on $\\pi_{\\le\nN}(X)\/\\Phi\\bigl(\\pi_{\\le N}(X)\\bigr)$. Then, G is nilpotent.\n\\end{theorem}\n\nIn particular, taking into account that for an abelian $p$-group\n$G$ its Frattini factor is precisely $G\\otimes\\mathbb Z\/p$, we obtain\nthe following:\n\n\\begin{corollary}\\label{nuevo} Assume that\n$\\pi_{\\le N}(X)$ is a finite abelian group and let $G$ be a\nsubgroup of $\\mathcal E(X)$ which acts nilpotently on $\\pi_{\\le\nN}(X)\\otimes\\mathbb Z\/p$ for any prime $p$. Then, G is\nnilpotent.\\hfill$\\square$\n\\end{corollary}\n\nNotice that, by the Universal Coefficients Theorem for homotopy,\n$\\pi_*X\\otimes\\mathbb Z\/p=\\text{Ext}(\\mathbb Z\/p,\\pi_*X)$ is a subgroup of\n$\\pi_*(X;\\mathbb Z\/p)$. Hence, as an immediate consequence of Corollary\n\\ref{nuevo} above we get:\n\n\\begin{corollary}\\label{coeficientes} Assume that\n$\\pi_{\\le N}(X)$ is a finite abelian group and let $G$ be a\nsubgroup of $\\mathcal E(X)$ which acts nilpotently on $\\pi_{\\le\nN}(X;\\mathbb Z\/p)$ for any prime $p$. Then, G is\nnilpotent.\\hfill$\\square$\n\\end{corollary}\n\n\n\n\nHaving studied the nilpotency of a general subgroup of $\\mathcal E(X)$,\nwe now focus on the group $\\mathcal E_{ \\sharp p }(X)$ and give necessary conditions\nfor it to be nilpotent.\n\n\\begin{theorem}\\label{tres} Let $X$ be a space for which $\\pi_{\\le N}(X)$ is a finite abelian $p$-group. Then\n$\\mathcal E_{ \\sharp p }(X)$ is nilpotent and $\\mathcal E_{ \\sharp p }(X)\/\\mathcal E_{ \\sharp }(X)$ is a finite $p$-group.\n\n\\end{theorem}\n\n\\begin{theorem}\\label{cuatro}\nLet $X$ be a space for which $\\pi_{\\le N}(X)$ is a finitely\ngenerated abelian group. Then $\\cap_{p\\,\\,\\text{prime}}\\mathcal E_{ \\sharp p }(X)$ is\nnilpotent.\n\\end{theorem}\n\n\\begin{remark}\\label{importante} {\\em Observe that in general\n$\\mathcal E_{ \\sharp p }(X)$ is bigger than $\\mathcal E_{ \\sharp }(X)$. For instance consider\n$X=K(\\mathbb Z\/{p^r},n)$, $r,n\\ge 2$. Obviously $\\mathcal E_{ \\sharp }(X)=\\{1\\}$, while\nthe automorphism $\\rho$ of $\\mathbb Z\/{p^r}$ given by\n$\\rho(1)=p^{r-1}+1$ induces a non trivial element of $\\mathcal E_{ \\sharp p }(X)$.\nIndeed, by the Universal Coefficients Theorem for homotopy, $$\n\\pi_*(X,\\mathbb Z\/p)=\\pi_n(X,\\mathbb Z\/p)\\oplus \\pi_{n-1}(X,\\mathbb Z\/p), $$ in\nwhich\\hfill\\break\n\n\\medskip\n\\noindent$\\pi_n(X,\\mathbb Z\/p)=\\hom(\\mathbb Z\/p,\\mathbb Z\/{p^r})$ {and}\n$\\pi_{n-1}(X,\\mathbb Z\/p)=\\text{Ext}(\\mathbb Z\/p,\\mathbb Z\/{p^r})=\\mathbb Z\/p.$\n\n\\medskip\n\\noindent Trivially $\\rho$ induces the identity on both. Note that\nthis example also shows that even\n$\\cap_{p\\,\\,\\text{prime}}\\mathcal E_{ \\sharp p }(X)$ can be bigger than $\\mathcal E_{ \\sharp }(X)$.}\n\\end{remark}\n\nThe paper is organized as follows: in the next section we collect\nthe results we shall need from group theory and from which\nTheorems \\ref{dos} and \\ref{uno} are immediately deduced. Theorem\n\\ref{local} and \\ref{principal} are proved in section \\S3.\nFinally, in section \\S4 we establish Theorems \\ref{tres} and\n\\ref{cuatro}.\n\n\\section{From group theory}\n\nWe begin by recalling some basic facts. If $G$ is a group acting\non another group $A$ (i.e., $A$ is a $G$-group), the $n$-th $G$-commutator subgroup\n$\\Gamma^n_G(A) \\subset A$ is the group generated by\n$\\{(ga^{-1})a\\,|\\, g\\in G,\\, a\\in\\Gamma_G^{n-1}(A)\\}$, being\n$\\Gamma_G^0(A)=A$. The action is then nilpotent of nilpotency\norder $r$, $\\text{nil}\\,_GA=r$, if this is the smallest integer for which\n$\\Gamma^r_G(A)=\\{1 \\}$. The group $G$ also acts in each\n$\\Gamma_G^n(A)$ and\n$\\Gamma_G^m\\bigl(\\Gamma_G^n(A)\\bigr)=\\Gamma_G^{m+n}(A)$.\n\nStatements of next sections shall heavily rely in the following\nresults:\n\n\\begin{lemma}\\label{lema1} Let $A$ be a $G$-group. Then:\n\\begin{itemize}\n\\item[(i)] $\\Gamma_G^1(A)$ is a normal subgroup of $A$ and the $G$-action induced on $A\/\\Gamma_G^1(A)$ is trivial.\n\\item[(ii)] The quotient morphism $A{\\stackrel{q}{\\longrightarrow}} A\/\\Gamma_G^1(A)$ is equivariant and initial with\nrespect to trivial actions, i.e., every equivariant morphism\n$A{\\stackrel{f}{\\longrightarrow}}H$, in which the $G$-action on\n$H$ is trivial, factors uniquely through $q$.\n\\end{itemize}\n\\end{lemma}\n\n\\begin{proof} (i) is trivial. For (ii) observe that, for any $f$ as in the lemma, $\\Gamma_G^1(A)\\subset\\ker f $.\n\\end{proof}\n\n\\begin{lemma}\\label{lema2} Let $A$ be a $G$-group. If $A$ is nilpotent then, for any $m$, $\\Gamma_G^m(A)_{(p)}=\\Gamma_G^m(A_{(p)})$.\n\\end{lemma}\n\n\\begin{proof} Since $\\Gamma_G^m(A)=\\Gamma_G^{1}\\bigl(\\Gamma_G^{m-1}(A)\\bigr)$, once we show that\n$\\Gamma_G^1(A)_{(p)}=\\Gamma^1_G(A_{(p)})$ an easy induction proves\nthe lemma. As localization is an exact functor in the category of\nnilpotent groups, the localization morphism $f:A\\to A_{(p)}$\nrestricts to $f:\\Gamma_G^1(A)\\to \\Gamma_G^1(A)_{(p)}$. Hence, we\nmay consider $\\Gamma_G^1(A)_{(p)}$, as well as\n$\\Gamma_G^1(A_{(p)})$, as subgroups of $A_{(p)}$. Then, for any\n$g\\in G$ and $a\\in A$, the trivial identity\n$\\bigl(gf(a)^{-1}\\bigr)f(a)=f\\bigl((ga^{-1})a\\bigr)$ shows\nequality of both subgroups.\n\\end{proof}\n\n\\begin{proposition}\\label{uf} The group $G$ acts nilpotently on the nilpotent group $A$ if and only if $G$ acts\nnilpotently on $A_{(p)}$ for $p$ any prime number or zero and all\nthese nilpotency orders are bounded.\n\\end{proposition}\n\n\\begin{proof} Assume $G$ acts nilpotently on $A$, i.e., $\\Gamma_G^m(A)=\\{1\\}$ for some $m$. Hence, by Lemma \\ref{lema2}\nand for any $p$, $\\Gamma_G^m(A_{(p)})=\\{1\\}$.\n\nConversely, assume $\\text{nil}\\,_GA_{(p)}\\le m$, for all $p$ ($p$ a prime\nnumber or $0$), and let $a$ be an element of $\\Gamma_G^m(A)$. If\n$a$ has finite order, say it is a $q$-element, then it obviously\nsurvives under the $q$-localization morphism\n$\\Gamma_G^m\\bigl(A)\\to\\Gamma^m_G(A)_{(q)}$. For a general group,\nelements of infinite order are not guaranteed to survive under\nrationalization (for instance, the rationalization of the free\nproduct of two finite groups is trivial while it contains elements\nof infinite order). However for a nilpotent group, which is our\ncase, one can easily show by induction on the nilpotency order of\nthe group, that any element of infinite order is not sent to zero\nunder rationalization. Taking into account, again by Lemma\n\\ref{lema2}, that $\\Gamma^m_G(A)_{(p)}=\\Gamma^m_G(A_{(p)})=\\{1\\}$,\nit follows that $a=1$ and the proof is complete.\n\\end{proof}\n\n\n\\begin{proposition}\\label{nuevolema} Let $G$ be a group acting on a finite nilpotent group $A$ in such a way that the\ninduced action on the Frattini factor $A\/\\Phi(A)$ is nilpotent.\nThen, the $G$-action on $A$ is also nilpotent.\n\\end{proposition}\n\n\\begin{proof} Recall \\cite[5.1]{Gorestein} that the Frattini subgroup of a group $A$, $\\Phi(A)$, is defined to be the intersection of all its maximal proper subgroups. The Frattini factor of $A$ is $A\/\\Phi(A)$.\nObserve in the first place that, since $\\Phi(A)$ is a characteristic subgroup of $A$, i.e., it is invariant under any automorphism of $A$, $G$ in fact induces a\nnatural action on the Frattini factor $A\/\\Phi(A)$ which, by\nhypothesis, is nilpotent. Hence, since $A\/\\Phi(A)$ is nilpotent,\nthe induced action on\n$\\bigl(A\/\\Phi(A)\\bigr)_{(p)}=A_{(p)}\/\\Phi(A)_{(p)}$ is also\nnilpotent by Lemma \\ref{lema2}. Next, observe that for any finite\ngroup $A$, $\\Phi(A)_{(p)}=\\Phi(A_{(p)})$. Indeed, this is\nimmediate from the definition taking into account that\nlocalization commutes with limits, in particular, with\nintersections (see for instance \\cite{hmr}). Therefore, we conclude that $G$ acts nilpotently on\n$A_{(p)}\/\\Phi(A_{(p)})$. Considering $\\varphi\\colon\nG\\to\\text{aut}\\,\\bigl(A_{(p)}\/\\Phi(A_{(p)})\\bigr)$ via this action, and\ntaking into account that $A_{(p)}\/\\Phi(A_{(p)})$ is a finite\n$p$-group, we may apply \\cite[Corollary 5.3.3]{Gorestein} to\nobtain that $\\varphi(G)$ is also a $p$-group. But the action of a\n$p$-group on another $p$-group is always nilpotent, and therefore\n$G$ acts nilpotently on $A_{(p)}$. Since this is the case for any\n$p$ and $A$ is finite we may apply Proposition \\ref{uf} and the\nproposition follows.\n\\end{proof}\n\nFrom these results we immediately deduce:\n\n\\bigskip\n\\noindent {\\it Proof of Theorems \\ref{dos} and \\ref{uno}.} Apply\ndirectly Propositions \\ref{uf} and \\ref{nuevolema} above to the\nsubgroup $G$ of $\\mathcal E (X)$ to obtain that $G$ acts nilpotently on\n$\\pi_{\\le N}(X)$. Then, the result follows from Theorem\n\\ref{principal}. \\hfill$\\square$\n\\bigskip\n\n\n\n\nClosely related to Proposition \\ref{nuevolema}, we have the\nfollowing:\n\n\\begin{proposition}\\label{propodos} Let $G$ be a group acting on an abelian $p$-group $A$ which has an exponent $p^n$. If\n$G$ acts nilpotently on $A\\otimes\\mathbb Z\/p$, then it does so on $A$\nand $$ \\text{nil}\\,_GA\\le n\\cdot\\text{nil}\\,_GA\\otimes\\mathbb Z\/p. $$\n\\end{proposition}\n\\begin{proof} Call $r=\\text{nil}\\,_GA\\otimes\\mathbb Z\/p$ and observe that $\\Gamma_G^m(A\\otimes\\mathbb Z\/p)=\\Gamma_G^m(A)\\otimes\\mathbb Z\/p$ for any\n$m$. Therefore, since $\\Gamma_G^r(A\\otimes\\mathbb Z\/p) =0$,\n$\\Gamma_G^r(A)\\subset pA$. Assume, as induction hypothesis, that\n$\\Gamma_G^{kr}(A)\\subset p^kA$, for $k0$. We then interpret the metric \\eqref{dsmetric} as modelling an expanding universe. \nIn this case the scale factor grows with time $t$ and the size of the horizon shrinks. \nWe note here that this observation stems from the convention that $t$ runs from $-\\infty $ to $+\\infty$ (which means conformal time $-\\infty<\\tau< 0$). Defining $\\bar{t}= -t$, which corresponds to \\say{$t$ running from $+\\infty$ to $-\\infty$ or $+\\infty>\\tau> 0$} we find the scale factor $a(\\tau)$ growing with $\\bar{t}$ for $H<0$, corresponding to expansion. Therefore, the expanding dS space can only characterized by shrinking horizon rather than fixing the notion of time being positive or negative. Since time is a parameter in quantum theory, we can have quantum states that are labeled by $\\tau<0$ and\/or $\\tau>0$. In the case of standard QFT in dS one usually make a biased choice of either $\\tau<0$ or $\\tau>0$ which actually breaks the time reversal nature of dS spacetime and also standard scheme of quantization completely ignores existence of states that evolve backward in time. \nTherefore, \n in our scheme of quantization in dS spacetime, we can have quantum states that propagate forward in time $t: -\\infty \\to \\infty$ with $H > 0$ and states that propagate backward in time following an arrow of time $t: \\infty \\to -\\infty$ with $H< 0$. Note that both type of quantum states can co-exist in a single expanding Universe. We stress again, the expanding dS Universe is defined by the following arrows of time\n\\begin{equation}\n\t\\tau: \\pm \\infty \\to 0\n\\end{equation}\nwhere $-$ sign corresponds to a state propagating forward in time and $+$sign corresponds to a state propagating backward in time. In our new picture quantization of a (massless) scalar field in dS spacetime leads to a pair of quantum fields that are produced out of two vacua. Being more precise we define total vacuum as the direct sum of two vacua\n\\begin{equation}\n\t\\vert 0\\rangle = \\frac{1}{\\sqrt{2}} \\Bigg(\\vert 0\\rangle_{\\rm I}\\oplus \\vert 0\\rangle_{\\rm II} \\Bigg)\\,. \n\\end{equation}\nIn the vacuum $\\vert 0\\rangle_{\\rm I}$ we create quantum fields at the position $ \\textbf{x} $ that evolve forward in time ($\\tau <0 $) and in the vacuum $\\vert 0\\rangle_{\\rm II}$ we create quantum fields at $ -\\textbf{x}$ which evolve backward in time ($\\tau >0$).\nThe action of a massless scalar field in dS space is given \n\\begin{equation}\nS_{\\phi} = -\\frac{1}{2}\\int d\\tau d^3x a^2\t\\phi\\left( \\partial_\\tau^2+2\\mathcal{H} \\partial_\\tau + k^2 \\right) \\phi\\,. \n\\end{equation}\nFirst thing we can notice here is that the above action is invariant under $\\mathcal{P}\\mathcal{T}$ which means $\\textbf{x}\\to -\\textbf{x}$ and \n\\begin{equation}\\label{timeref}\n\tt\\to -t\\quad \\left(-\\tau \\to \\tau\\right) \\implies H\\to -H \\quad \\left( \\mathcal{H}\\to -\\mathcal{H} \\right)\n\\end{equation}\nRescaling the field $\\phi\\to a \\phi$ gives the following action which leads to Mukhanov-Sasaki equation as\n\\begin{equation}\n\tS_\\phi = \\frac{1}{2} \\int d\\tau d^3x \\Big[ \\phi^{\\prime 2} - \\left( \\partial_i\\phi \\right)^2 - \\frac{2}{\\tau^2} \\phi^2 \\Big]\\,. \n\t\\label{msaction}\n\\end{equation}\nThe above Lagrangian is again symmetric under $\\mathcal{P}\\mathcal{T}: \\textbf{x}\\to -\\textbf{x},\\, \\tau\\to -\\tau$. When we quantize this scalar field\nwe propose that the scalar field operator $\\hat{\\phi}\\left( \\tau,\\, \\textbf{x} \\right)$ is the direct sum of the pair of fields as\n\\begin{equation}\n\t\\hat{\\phi}\\left( \\tau,\\, \\textbf{x} \\right) = \\frac{1}{\\sqrt{2}}\t\\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\oplus \t\\frac{1}{\\sqrt{2}}\\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right) \\,. \n\\end{equation}\nwhich can be expanded in terms of different creation and annihilation operators in the following way\n\\begin{equation}\n\t\\begin{aligned}\n\t&\t\\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\\\ & = \\frac{1}{\\left( 2\\pi \\right)^{3\/2}}\\int d\\tau d^3k\\Bigg[ a_\\textbf{k} {\\varphi}_{\\rm I\\,k}\\left( \\tau \\right) e^{i\\textbf{k}\\cdot \\textbf{x}} + a_\\textbf{k}^\\dagger {\\varphi}^\\ast_{\\rm I\\,k}\\left( \\tau \\right) e^{-i\\textbf{k}\\cdot \\textbf{x}} \\Bigg]\\, \n\t\\end{aligned}\n\\label{vdfield}\n\\end{equation}\nwhere the creation and annihilation operators $a_\\textbf{k},\\, a^\\dagger_\\textbf{k}$ satisfy the canonical commutation relations and \n \\begin{equation}\na_\\textbf{k}\\vert 0 \\rangle_{\\rm I} = 0\\,. \n\\end{equation}\nHere the mode function $\\varphi_{\\rm I\\,,k}$ is the solution of Mukhanov-Sasaki equation (that comes from varying \\eqref{msaction}) given by \n\\begin{equation}\n\t\\varphi_{\\rm I,\\,k} = \\alpha_{1k} \\frac{e^{-ik\\tau}}{\\sqrt{2k}}\\left( 1-\\frac{i}{k\\tau} \\right) +\\beta_{1k} \\frac{e^{ik\\tau}}{\\sqrt{2k}}\\left( 1+\\frac{i}{k\\tau} \\right)\\,, \n\t\\label{f1}\n\\end{equation}\nwhere the Bogoliubov coefficients can be fixed as $\\left( \\alpha_{1,\\,k},\\, \\beta_{1,\\,k}\\right) = \\left( 1,\\,0 \\right)$, which is compatible with the Wronskian condition $\t\\varphi_{\\rm I,\\,k} \t\\varphi^{\\prime\\ast}_{\\rm I,\\,k}-\t\\varphi^\\ast_{\\rm I,\\,k}\t\\varphi^\\prime_{\\rm I,\\,k} = i $ that corresponds to the canonical commutation relation\n\\begin{equation} \n\\big[ \\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right),\\, \\hat{\\pi}_{\\rm I}\\left( \\tau,\\, \\textbf{x}^\\prime \\right)\\big] = i\\delta\\left( \\textbf{x}-\\textbf{x}^\\prime \\right). \n\\end{equation}\nThe choice $\\left( \\alpha_{\\rm I,\\,k},\\, \\beta_{\\rm I,\\,k}\\right) = \\left( 1,\\,0 \\right)$ defines the vacuum and $\\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\vert 0\\rangle $ corresponds to the positive frequency modes in the limit $\\tau\\to -\\infty$. Quantum mechanically $\\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\vert 0\\rangle $ is the postive energy state that propagate forward in time at \\textbf{x}. \nSimilarly, we can expand the second field operator as \n\\begin{equation}\n\t\\begin{aligned}\n\t\t&\t\\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right) \\\\ & = \\frac{1}{\\left( 2\\pi \\right)^{3\/2}}\\int d\\tau d^3k\\Bigg[ b_\\textbf{k} \\varphi_{\\rm II\\,k}\\left( -\\tau \\right) e^{-i\\textbf{k}\\cdot \\textbf{x}} + b_\\textbf{k}^\\dagger {\\varphi}^\\ast_{\\rm II\\,k}\\left( -\\tau \\right) e^{i\\textbf{k}\\cdot \\textbf{x}} \\Bigg]\\, \n\t\\end{aligned}\n\t\\label{vdfield2}\n\\end{equation}\nwhere \n\\begin{equation}\n\t\\varphi_{\\rm II,\\,k} = \\alpha_{2,\\,k} \\frac{e^{ik\\tau}}{\\sqrt{2k}}\\left( 1+\\frac{i}{k\\tau} \\right) +\\beta_{2,\\,k} \\frac{e^{-ik\\tau}}{\\sqrt{2k}}\\left( 1-\\frac{i}{k\\tau} \\right)\\,, \n\t\\label{f2}\n\\end{equation}\nwhere the Bogoliubov coefficients can be fixed as $\\left( \\alpha_{2,\\,k},\\, \\beta_{2,\\,k}\\right) = \\left( 1,\\,0 \\right)$ which is compatible with the Wronskian condition $\t\\varphi_{\\rm II,\\,k} \t\\varphi^{\\prime\\ast}_{\\rm II,\\,k}-\t\\varphi^\\ast_{\\rm II,\\,k}\t\\varphi^\\prime_{\\rm II,\\,k} = -i $ that corresponds to the canonical commutation relation \\cite{Donoghue:2019ecz}\n\\begin{equation} \n\t\\big[ \\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right),\\, \\hat{\\pi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x}^\\prime \\right)\\big] = -i\\delta\\left( \\textbf{x}-\\textbf{x}^\\prime \\right). \n\\end{equation}\nwhich describe the quantum fields that propagate backward in time. \nHere the second vacuum is defined by \n\\begin{equation}\n\tb_\\textbf{k}\\vert 0 \\rangle_{\\rm II} = 0\\,. \n\\end{equation}\nHere $\t\\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right)\\vert 0\\rangle_{\\rm II}$ corresponds to positive energy state that evolves backward in time at -\\textbf{x}. \nWe demand that these two quantum fluctuations evolve independently. This is manifest by\n\\begin{equation}\n\t[\t\\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right),\\, \t\\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right) ] =0\\,. \n\\end{equation}\nwhich implies that the respective creation and annihilation operators commute\n\\begin{equation}\n\t\\big[a_\\textbf{k},\\, b_{\\textbf{k}^\\prime}\\big] = 0,\\quad \\big[a^\\dagger_\\textbf{k},\\, b^\\dagger_{\\textbf{k}^\\prime}\\big] = 0 \n\\end{equation}\nSince dS spacetime is perfectly $\\mathcal{P}\\mathcal{T}$ symmetric we have that the quantum fields $\t\\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right)\\vert 0\\rangle_{\\rm I}$ and $\t\\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right)\\vert 0\\rangle_{\\rm II}$ behave identically, which can be seen from the fact that their equal time correlations are the same\n\\begin{equation}\n\t\\begin{aligned}\n& \\frac{1}{a^2}{}_{\\rm I}\\langle 0\\vert \\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\hat{\\varphi}_{\\rm I}\\left( \\tau,\\, \\textbf{x}^\\prime \\right)\\vert 0\\rangle_{\\rm I} = \\\\ & \\frac{1}{a^2} {}_{\\rm II}\\langle 0\\vert \\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right) \\hat{\\varphi}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x}^\\prime \\right)\\vert 0\\rangle_{\\rm II} = \\frac{H^2}{4\\pi^2k^3}\\,. \n\t\\end{aligned}\n\\label{eqcorr}\n\\end{equation}\nThis implies the correlations of quantum fields related by $\\mathcal{P}\\mathcal{T}$ transformations are identical in the case of dS spacetime. \nThis indicates our formulation of QFT in dS spacetime might be $\\mathcal{P}\\mathcal{T}$ or $\\mathcal{C}\\mathcal{P}\\mathcal{T}$ (if we include charged fields) symmetric.\\footnote{We note that the result in \\eqref{eqcorr} is similar to what was derived in the context of elliptic dS spacetime \\cite{SANCHEZ19871111,Schrodinger1956}. However, in our scheme of quantization we do not make any (classical) antipodal identification of spacetime and therefore we significantly deviate conceptually from the proposal of elliptic dS made by Erwin Schr\\\"{o}dinger \\cite{Schrodinger1956}. }\n\nHowever, this may not be true in a spacetime with less symmetries such as quasi-dS. Indeed, in the context of inflationary quantum fluctuations, it is often assumed, in the effective field theory (EFT) approach, that Lorentz symmetry (in other words $(\\mathcal{C})\\mathcal{P}\\mathcal{T}$) must be spontaneously broken \\cite{Cheung:2007st}. However, it is not known how $\\mathcal{C}\\mathcal{P}\\mathcal{T}$ symmetry can be broken in a curved spacetime. \n\t In the next section, we provide a framework of quantization where we can explicitly see that $(\\mathcal{C})\\mathcal{P}\\mathcal{T}$ is spontaneously broken in the sense that the qunatum fields defined by our new notion of $\\mathcal{P}\\mathcal{T}$ transformations evolve differently. \n\n\\section{Double vacuum inflationary spacetime and the power spectra}\nIn the previous section, we have studied the quantization of a massless scalar field in dS. Here we quantize the inflationary quantum fluctuations. There is a crucial difference here. In the previous section we quantized an arbitrary scalar field in dS spacetime which is completely fixed but in the context of inflation we quantize metric and matter degrees of freedom in order to find an effective quantum correction to the classical quasi-dS (inflationary) spacetime. To be more precise, the classical inflationary spacetime is given by an arrow of time and the background initial conditions are \n\\begin{equation}\nt: -\\infty \\to \\infty \\implies \\left( \\epsilon,\\, \\eta \\right): \\left( \\approx 0,\\, \\approx 0 \\right) \\to \\left( \\sim 1,\\, \\sim 1 \\right)\\,. \n\\end{equation}\nWhich means the slow-roll parameters are very small and positive during inflation and inflation ends when the slow-roll parameters evolve to the order of unity, which happens at some time $t_{\\rm end}\\gg 1$. The inflationary quantum fluctuations are produced in the phase when $\\left( \\epsilon,\\, \\eta \\right) \\ll 1$. As we discussed earlier, ``time'' in quantum theory has totally different role and one must be very careful in quantizing inflationary fluctuations. We emphasize that the quantum fluctuations do not necessarily follow arrow of time prescribed by the initial conditions in classical theory. \n\n Unlike dS spacetime the quasi-dS metric does not have $\\mathcal{P}\\mathcal{T}$-symmetry, therefore naturally one would expect the $\\mathcal{P}\\mathcal{T}$-symmetry to be spontaneously broken at the quantum level. To see this, we apply the scheme of quantization similar to what we have done in the context of exact dS. Let us first focus on inflationary scalar fluctuations which are described by the action \\eqref{scalar}. After the field redefinition $\\hat{v}$ is the canonical variable which we need to quantize \\eqref{screfa}. As we postulated before quantum fluctuations always come as pair from two different vacua related by $\\mathcal{P}\\mathcal{T}$ transformations. This means that unlike the standard quantization of inflationary fluctuations, where only one kind of quantum states are considered which only evolve forward in time, here we describe a pair of quantum fluctuations that evolve both forward and backward in time. \n\n Similar to the exact dS case we write the canonical field operator as a direct sum of two quantum fields \n\\begin{equation}\n\t\\hat{v}\\left( \\tau,\\, \\textbf{x} \\right) = \\frac{1}{\\sqrt{2}} \\hat{v}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\oplus \\frac{1}{\\sqrt{2}} \\hat{v}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right)\\,. \n\\end{equation}\nwhich emerges from two different vacua. The total vacuum is the direct sum of these two which can be written as\n\\begin{equation}\n\t\\vert 0\\rangle_{\\rm qdS} = \\vert 0 \\rangle_{\\rm qdS_{I}} \\oplus \\vert 0\\rangle_{\\rm qdS_{II}}\\,. \n\\end{equation}\nThe quantum field operators acting on the vacua $\\hat{v}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\vert 0\\rangle_{\\rm qdS_{II}} $ and $\\hat{v}_{\\rm I}\\left( -\\tau,\\, -\\textbf{x} \\right)\\vert 0\\rangle_{\\rm qdS_{II}}$ leads to the creation of the pair of quantum fluctuations at $ \\textbf{x} $ and at $ -\\textbf{x} $ which propagate forward in time and back ward in time respectively. We will make it clear shortly the meaning of quantum fluctuations that evolve forward and backward in time in the context of quasi-dS.\n\nGoing into the details\nwe expand the fields in terms of creation and annhilation operators as explained below. \n\\begin{equation}\n\t\\begin{aligned}\n& \\hat{v}_{\\rm I}\\left( \\tau,\\, \\textbf{x} \\right) \\\\ & =\t \\frac{1}{\\left( 2\\pi \\right)^{3\/2}}\\int d\\tau d^3k\\Bigg[ c_\\textbf{k} {v}_{\\rm I,\\,k}\\left( \\tau \\right) e^{i\\textbf{k}\\cdot \\textbf{x}} + c_\\textbf{k}^\\dagger {v}_{\\rm I,\\,k}^\\ast\\left( \\tau \\right) e^{-i\\textbf{k}\\cdot \\textbf{x}} \\Bigg]\n\\end{aligned}\n\\label{vid}\n\\end{equation}\nwith $c_\\textbf{k},\\, c_\\textbf{k}^\\dagger$ being the creation and annihilation operators of quasi-dS vacuum-I defined by\n\\begin{equation}\n\tc_\\textbf{k}\\vert 0\\rangle_{\\rm qdS_I} = 0\\,, \n\t\\label{qds1}\n\\end{equation}\nand $ {v}_{\\rm I,\\,k}$ is the mode function obtained by solving MS-equation for $v_{\\rm I}\\left(\\tau \\right)$ \n\\begin{equation}\n\tv_{\\rm I,\\,k}^{\\prime\\prime}+ \\left( k^2-\\frac{{\\nu}_s^2-\\frac{1}{4}}{\\tau^2} \\right) v_{\\rm I,\\,k}^2 =0\\,.\n\\end{equation}\nHere $\\tau: -\\infty \\to 0$ and \n the quasi-dS vacuum $\\vert 0\\rangle_{\\rm qdS_I}$ corresponds to the creation of positive frequency modes in the limit $\\tau\\to -\\infty$. \nNotice that unlike the dS case the MS-equation \\eqref{MS-equation} is not symmetric under time reversal because there is an additional time dependence enters through the nearly constant variable $\\nu_s$ which contains the slow-roll parameters $\\left( \\epsilon,\\, \\eta \\right)$ \\eqref{sldef}. Assuming $0<\\left( \\epsilon,\\, \\eta\\right) \\ll 1$ and constant we obtain \n\\begin{equation}\n\t\\begin{aligned}\n&\t{v}_{\\rm I,\\,k} \\\\ & = \\frac{\\sqrt{\\pi}}{2} \\left( -\\tau \\right)^{1\/2} e^{\\left( i\\nu_s+1\\right)} \\Bigg[C_k H^{(1)}_{\\nu_s}\\left( -k \\tau \\right)+ D_k H^{(2)}_{\\nu_s}\\left( -k \\tau \\right)\\Bigg]\\,. \n\t\\label{new-vac1}\n\t\\end{aligned}\n\\end{equation}\nThe above mode function corresponds to the creation of positive frequency modes in the limit $\\tau\\to -\\infty$ for the case $\\left( \\mathcal{C}_k,\\, D_k \\right) = \\left( 1,\\,0 \\right) $ which corresponds to the standard BD state that satisfies the Wronskian $v_{I,k}v_{I,k}^{\\prime\\ast}-v_{I,k}^\\ast v_{I,k}^{\\prime}=i (\\implies \\vert C_k\\vert^2-\\vert D_k\\vert^2=1)$. We can expand the mode function for the case of $\\left( \\mathcal{C}_k,\\, D_k \\right) = \\left( 1,\\,0 \\right) $ in the order of slow-roll as\n\\begin{equation}\n\t\\begin{aligned}\n\t{v}_{\\rm I,\\,k} & \\approx \\frac{\\sqrt{\\pi}}{2} \\left( -\\tau \\right)^{1\/2} e^{\\left( i\\nu_s+1\\right)} H^{(1)}_{\\nu_s}\\left( -k \\tau \\right)\\,\\\\ \n\t& \\approx \\sqrt{\\frac{1}{2k}} e^{-ik\\tau}\\left( 1-\\frac{i}{k\\tau} \\right) \\\\ & + \\left( \\epsilon+\\frac{\\eta}{2} \\right) \\frac{\\sqrt{\\pi}}{2\\sqrt{k}} \\sqrt{\\frac{k}{aH}} \\frac{\\partial H^{(1)}_{\\nu_s}\\left( \\frac{k}{aH}\\right)}{\\partial \\nu_s}\\Big\\vert_{\\nu_s=3\/2}\n\\label{new-vac1BD}\n\\end{aligned}\n\\end{equation}\n\nThe second field $\\hat{v}_{II\\,,\\textbf{k}}\\left( -\\tau,\\, -\\textbf{x} \\right)$ is expanded as \n\\begin{equation}\n\t\\begin{aligned}\n&\t\\hat{v}_{\\rm II}\\left( -\\tau,\\, -\\textbf{x} \\right) \\\\ & =\t \\frac{1}{\\left( 2\\pi \\right)^{3\/2}}\\int d\\tau d^3k\\Bigg[ d_\\textbf{k} {v}_{\\rm II,\\,k}\\left( -\\tau \\right) e^{-i\\textbf{k}\\cdot \\textbf{x}} + d_\\textbf{k}^\\dagger {v}_{\\rm I,\\,k}^\\ast\\left( -\\tau \\right) e^{i\\textbf{k}\\cdot \\textbf{x}} \\Bigg]\n\t\\end{aligned}\n\\label{v2d}\n\\end{equation}\nwhere $d_\\textbf{k},\\ , d_\\textbf{k}^\\dagger$ are the creation and annihilation operators that satisfy the canonical commutation relations and the second vacuum is defined by\\footnote{We note that $\n\t\t\\big[c_\\textbf{k},\\, d_{\\textbf{k}^\\prime}\\big] = 0,\\quad \\big[c^\\dagger_\\textbf{k},\\, d^\\dagger_{\\textbf{k}^\\prime}\\big] = 0 $ similar to the dS case which means the pair of quantum fluctuations are not causally connected.}\n\\begin{equation}\nd_\\textbf{k}\\vert 0\\rangle_{\\rm qdS_{II}}=0\\,.\n\\end{equation}\nThe evolution of the mode function $v_{\\rm II,\\,k}$ is determined by solving the following time reversed MS-equation \n\\begin{equation}\n\tv_{\\rm II,\\,k}^{\\prime\\prime}+ \\left( k^2-\\frac{\\bar{\\nu}_s^2-\\frac{1}{4}}{\\tau^2} \\right) v_{\\rm II,\\,k}^2 =0\\,.\n\\end{equation}\nHere $\\tau: +\\infty \\to 0$ and\n\\begin{equation}\n\t\\bar{\\nu}_s \\approx \\frac{3}{2}-\\epsilon-\\frac{\\eta}{2}\n\\end{equation}\nHere the time reversal transformations are \n\\begin{equation}\n\tt\\to -t \\implies H\\to -H,\\quad \\epsilon\\to -\\epsilon,\\quad \\eta\\to -\\eta\\,. \n\t\\label{timerevqds}\n\\end{equation}\nwhich can be understood in the following sense. Our goal here is to define how the quantum fluctuations propagate backward in time in an expanding Universe. To achieve this goal we need to formulate the meaning of time reversal operation. Logically, if the fluctuation propagates forward in time in a slow-roll background, the fluctuation that goes backward in time experience spacetime as a \"slow-climb\"\\footnote{In the standard slow-roll we have scalar field rolling down the potential whose time reversal can be understood as a phantom field that is slowly climbing the potential \\cite{Piao:2004tq}. } which is given by reversing the signs of the parameters $\\left( \\epsilon,\\, \\eta \\right)$ as stated in \\eqref{timerevqds}. We impose this time reversal operation in a completely quantum mechanical sense and this has no classical meaning i.e., our background (classical) dynamics are completely determined by Friedmann equations \\eqref{Freq} and we do not apply at all time reversal to the classical background. Since we treat time differently at quantum level we restrain ourselves from any intuition from classical physics.\\footnote{Notion of time is a very non-trivial concept in physics and its meaning varies in different contexts. We suggest the reader \\cite{Rovelli:2004tv} for an extended physical discussion. Our statement, that quantum states evolving backward in time has no classical analog, is deeply rooted in the quantum gravity concept time as it is presented in p. 184 of \\cite{Rovelli:2004tv}.} Since dynamics of quantum fields emerge from MS-equation \\eqref{MS-equation} the functions $\\left( \\epsilon,\\, \\eta \\right)$ are now treated along with time in as parameters to specify the nature of quantum state. This would encode a subtle difference between quantum fluctuations propagating forward and backward in time.\n\n\\begin{equation}\n\t\\begin{aligned}\n\t\t\t&\tv_{\\rm II,\\,k}\\left( -\\tau \\right) \\\\ &= \\dfrac{\\sqrt{\\pi}}{2} \\left( \\tau \\right)^{1\/2} e^{i\\left( 2\\bar{\\nu}_s+1\\right) \\pi}\\Bigg[\\bar{C}_k H^{(1)}_{\\bar{\\nu}_s}\\left( k \\tau \\right)+ \\bar{D}_k H^{(2)}_{\\bar{\\nu}_s}\\left( k \\tau \\right)\\Bigg].\n\t\t\t\\end{aligned}\n\t\t\\label{v2sol}\n\\end{equation}\nNow imposing the Wronskian condition $v_{\\rm II,\\,k}v_{\\rm II,\\,k}^{\\prime\\ast}-v_{\\rm II,\\,k}^\\ast v_{\\rm II,\\,k}^\\prime = -i$ which corresponds to the canonical commutation relation for a reversed arrow of time \\cite{Donoghue:2019ecz}\n\\begin{equation}\n\t\\Big[ \\hat{v}_{\\rm II}\\left( -\\tau,\\,-\\textbf{x} \\right),\\, \\hat{\\Pi}_{\\rm II}\\left( -\\tau,\\,-\\textbf{x}^\\prime \\right) \\Big] = -i \\delta\\left( \\textbf{x}-\\textbf{x}^\\prime \\right)\\,. \n\\end{equation}\nwe obtain $\\vert \\bar{C}_k\\vert^2-\\vert \\bar{D}_k\\vert^2 = 1$. We consider $\\left(\\bar{C},\\, \\bar{D}_k\\right) =0$ which corresponds postive energy states in the limit $\\tau \\to \\infty$. Expanding \\eqref{v2sol} up to the leading order for $\\left(\\bar{C},\\, \\bar{D}_k\\right) =0$ we get\n\\begin{equation}\n\t\\begin{aligned}\n\t\t{v}_{\\rm II,\\,k} & \\approx \\frac{\\sqrt{\\pi}}{2} \\left( \\tau \\right)^{1\/2} e^{\\left( i\\nu_s+1\\right)} H^{(1)}_{\\nu_s}\\left( k \\tau \\right)\\,\\\\ \n\t\t& \\approx \\sqrt{\\frac{1}{2k}} e^{ik\\tau}\\left( 1+\\frac{i}{k\\tau} \\right) \\\\ & -\\left( \\epsilon+\\frac{\\eta}{2} \\right) \\frac{\\sqrt{\\pi}}{2\\sqrt{k}} \\sqrt{k\\tau} \\frac{\\partial H^{(1)}_{\\nu_s}\\left( k\\tau \\right)}{\\partial \\nu_s}\\Big\\vert_{\\nu_s=3\/2}\n\t\t\\label{new-vac2BD}\n\t\\end{aligned}\n\\end{equation}\nComparing \\eqref{new-vac1BD} and \\eqref{new-vac2BD} we can deduce that the two mode functions get different slow-roll (quantum) corrections. \nThis is because the parameter $\\nu_s$ that enters in the MS-equation \\eqref{MS-equation} contains a tiny time dependence. In the standard framework of quantization this is treated it to be nearly constant there is only one type of quantum fluctuation (see Sec.~\\ref{sec:review}). But the fact that MS-equation has time dependence in terms of $\\epsilon,\\, \\eta$ is crucial to probe the nature of quantum fluctuations. These quantities change sign under discrete spacetime transformation and this has to be understood in a completely quantum mechanical sense. Our scheme of quantization implies a pair of quantum fluctuations: those that propagate forward in time ($\\tau: -\\infty \\to 0$) at spatial position $\\textbf{x}$ which behave differently from those fluctuations that propagate backward in time ($\\tau: +\\infty \\to 0$) at the spatial position $-\\textbf{x}$. These pairs of fluctuations exit the horizon in the two opposite directions and now we observe the modes as they re-enter. \n\nSo far we applied our quantization method for the scalar mode $v\\left( \\tau,\\,\\textbf{x} \\right)$. We can apply the similar method of quantization for the tensor modes \\eqref{tensor-mode} by writing the tensor fluctuation as direct some fluctuations that are related by $\\mathcal{P}\\mathcal{T}$ transformations \n\\begin{equation}\n\t\\hat{u}_{ij} \\left( \\tau,\\, \\textbf{x} \\right)= \\frac{1}{\\sqrt{2}} \\hat{u}^{\\rm I}_{ij}\\left( \\tau,\\, \\textbf{x} \\right) \\oplus \\frac{1}{\\sqrt{2}}\\hat{u}^{\\rm II}_{ij}\\left( -\\tau,\\,-\\textbf{x} \\right)\n\\end{equation}\nThe above fields can be written in terms of creation and annihilation operators similar to what we have done for \\eqref{vid} and \\eqref{v2d}. The corresponding mode functions tensor modes can be straight forwardly derived as\n\\begin{equation}\n\t\\begin{aligned}\n\t\t{u}^{\\rm I}_{ij,\\,k} \n\t\t& \\approx e_{ij}\\sqrt{\\frac{1}{2k}} e^{-ik\\tau}\\left( 1-\\frac{i}{k\\tau} \\right) \\\\ & + e_{ij} \\epsilon\\frac{\\sqrt{\\pi}}{2\\sqrt{k}} \\sqrt{-k\\tau} \\frac{\\partial H^{(1)}_{\\nu_t}\\left( -k\\tau \\right)}{\\partial \\nu_t}\\Big\\vert_{\\nu_t=3\/2}\n\t\\end{aligned}\n\t\t\\label{new-vac1TBD}\n\\end{equation}\nand \n\\begin{equation}\n\t\\begin{aligned}\n\t\t{u}^{\\rm II}_{ij,\\,k} \n\t\t& \\approx e_{ij}\\sqrt{\\frac{1}{2k}} e^{ik\\tau}\\left( 1+\\frac{i}{k\\tau} \\right) \\\\ & - e_{ij} \\epsilon\\frac{\\sqrt{\\pi}}{2\\sqrt{k}} \\sqrt{k\\tau} \\frac{\\partial H^{(1)}_{\\nu_t}\\left( k\\tau \\right)}{\\partial \\nu_t}\\Big\\vert_{\\nu_t=3\/2}\n\t\\end{aligned}\n\t\t\\label{new-vac2TBD}\n\\end{equation}\nwhere $e_{ij}$ denotes the polarization tensor. \nIn the next section, we will compute the inflationary power spectra for scalar and tensor modes.\n\n\\section{Inflationary power spectra and predictions for hemispherical asymmetry}\n\nAs we learned in the previous section, inflationary quantum fluctuations now come in pair which exit the horizon on two opposite sides. The two point correlations of these fluctuations can be computed as \n\\begin{equation}\n\t\\begin{aligned}\n\t\t&\t{}_{\\rm qdS_{I}}\\langle 0 \\vert \\hat{v}_{\\rm I}\\left( \\textbf{x},\\, \\tau \\right) \\hat{v}_{\\rm I}\\left( \\textbf{x}+\\boldsymbol{\\xi}, \\, \\tau \\right)\\vert 0\\rangle_{\\rm qdS_{I}} = \\\\ & \\frac{4\\pi}{\\left( 2\\pi \\right)^3} \\int \\frac{dk}{k} \\frac{\\sin k \\xi}{k\\xi} k^3 \\vert v_{\\rm I,\\,k}\\vert^2\\, \\\\ \n\t\t&\t{}_{\\rm qdS_{II}}\\langle 0 \\vert \\hat{v}_{\\rm II}\\left( -\\textbf{x},\\, -\\tau \\right) \\hat{v}_{\\rm II}\\left( -\\textbf{x}-\\boldsymbol{\\xi}, \\, -\\tau \\right)\\vert 0\\rangle_{\\rm qdS_{II}} = \\\\ &\\frac{4\\pi}{\\left( 2\\pi \\right)^3} \\int \\frac{dk}{k} \\frac{\\sin k \\xi}{k\\xi} k^3 \\vert v_{\\rm II,\\,k}\\vert^2\\,. \n\t\\end{aligned}\n\\label{power-spectrav}\n\\end{equation}\nSubstituting the expressions \\eqref{new-vac1BD} and \\eqref{new-vac2TBD} in the above expressions we learn that the above two correlations are not equal. In the case of exact dS we obtained the correlations to be equal (see \\eqref{eqcorr}) but in the case of quasi-dS we obtained an asymmetry because the background spacetime is not $\\mathcal{P}\\mathcal{T}$ symmetric. \nAs we know that the curvature perturbation is frozen on super-horizon scales. \nTo calculate the two point correlations of curvature perturbation on super-horizon scales we re scale the canonical fields with the classical background quantities \\eqref{rescaleeq}. This implies \n\\begin{equation}\n\t\\begin{aligned}\n{}_{\\rm qdS}\\langle 0 \\vert \\zeta_\\textbf{k} \\zeta_{\\textbf{k}^\\prime} \\vert 0\\rangle_{\\rm qdS} & = \\left( \\frac{1}{2a^2\\epsilon}\\right)\\Bigg\\vert_{\\rm classical} \\Big[\\frac{1}{2}{}_{\\rm qdS_I}\\langle 0 \\vert \\hat{v}_{\\rm I\\,\\textbf{k}} \\hat{v}_{\\rm I\\,\\textbf{k}^\\prime} \\vert 0\\rangle_{\\rm qdS_I}\\\\ & \\quad +\\frac{1}{2}{}_{\\rm qdS_{II}}\\langle 0 \\vert \\hat{v}_{\\rm {II}\\,\\textbf{k}} \\hat{v}_{\\rm {II}\\,\\textbf{k}^\\prime} \\vert 0\\rangle_{\\rm qdS_{II}}\\Big] \\\\ \n& = \\frac{2\\pi^2}{k^3}\\left( P_{\\zeta_1}+P_{\\zeta_2} \\right) \\delta\\left( \\textbf{k}+\\textbf{k}^\\prime \\right)\\,, \n\\end{aligned}\n\\end{equation}\nwhere $\\zeta_1,\\,\\zeta_2$ are the curvature perturbations that exit the horizon on two opposite directions and they become frozen on super-horizon scales which follows from \\eqref{power-spectrav}. With appropriate normalization we evaluate the power spectrum of these curvature perturbations at the moment of horizon exit. Computing the two power spectra at the horizon exit we obtain\n\\begin{equation}\n\t\\begin{aligned}\n\tP_{\\zeta1} & = \\frac{k^3}{2\\pi^2}\\frac{1}{2a^2\\epsilon} \\vert v_{\\rm I,\\,k}\\vert^2\\Bigg\\vert_{\\tau = -\\frac{1}{aH}} \\\\ \n\t& \\approx \\frac{H^2}{8\\pi^2\\epsilon} \\left( 1+\\left( \\frac{k}{aH} \\right)^2 \\right) \\\\ &\\quad + \\frac{H^2}{4\\pi\\epsilon} \\left( \\epsilon+\\frac{\\eta}{2} \\right) \\left(\\frac{k}{aH}\\right)^3 H_{3\/2}^{(1)} \\left( \\frac{k}{aH} \\right) \\frac{\\partial H^{(1)}_{\\nu_s}\\left( \\frac{k}{aH} \\right)}{\\partial \\nu_s}\\Big\\vert_{\\nu_s=3\/2}\\, \\\\\n\t\tP_{\\zeta2} & = \\frac{k^3}{2\\pi^2}\\frac{1}{2a^2\\epsilon} \\vert v_{\\rm II,\\,k}\\vert^2\\Bigg\\vert_{\\tau_\\ast = \\frac{1}{aH}} \\\\ \n\t& \\approx \\frac{H^2}{8\\pi^2\\epsilon} \\left( 1+\\left( \\frac{k}{aH} \\right)^2 \\right) \\\\ & - \\frac{H^2}{4\\pi\\epsilon} \\left( \\epsilon+\\frac{\\eta}{2} \\right) \\left(\\frac{k}{aH}\\right)^3 H_{3\/2}^{(1)} \\left( \\frac{k}{aH} \\right) \\frac{\\partial H^{(1)}_{\\nu_s}\\left( \\frac{k}{aH} \\right)}{\\partial \\nu_s}\\Big\\vert_{\\nu_s=3\/2}\\,. \n\t\\end{aligned}\n\\label{pw12}\n\\end{equation}\nIf we combine the above two expressions we obtain \n\\begin{equation}\n\tP_\\zeta = \\frac{1}{2}P_{\\zeta1}+ \\frac{1}{2} P_{\\zeta2} = \\frac{H^2}{8\\pi^2\\epsilon}\\Bigg\\vert_{k=aH}\n\\end{equation}\nwhich is the expression in the context of standard inflation (see Sec.~\\ref{sec:review}). From \\eqref{power-spectrav} we can deduce that the power spectrum $P_{\\zeta 1}$ can be mapped to the two-point correlations in the direction $\\hat{\\textbf{n}}$ and $P_{\\zeta 2}$ can be mapped to the two-point correlations in the opposite direction $-\\hat{\\textbf{n}}$ of the CMB sky. The difference between the two power spectra gives us the non-zero scale dependent contribution. Our formalism of producing inflationary correlations in pair naturally explains the hemispherical asymmetry and this can also be interpreted as an indication of spontaneous breaking of $\\mathcal{P}\\mathcal{T}$ symmetry in the inflationary background. In our context we can define the \nthe amplitude of dipolar modulation as (according to Eq.~\\eqref{Akg})\n\\begin{equation}\n\tA(k) = \\frac{P_{\\zeta1}-P_{\\zeta2}}{4P_\\zeta}\n\t\\label{Ako}\n\\end{equation}\nFrom \\eqref{pw12} we can notice that the two power spectra differ only by a small scale dependent correction of the order of slow-roll parameters. In addition, through CMB we can only probe very limited range of $k$ corresponding to initial $7-8$ e-foldings centered around the pivot scale \\cite{Martin:2004um}. \n At the leading order the first terms in the two power spectra dominate which gives us the tilt of the two power spectra nearly the same in the leading order in slow-roll approximation \n\\begin{equation}\n\t\\frac{d\\ln P_{\\zeta 1}}{d\\ln k}\\approx \t\\frac{d\\ln P_{\\zeta 2}}{d\\ln k} \\approx n_s-1 \\approx -2\\epsilon-\\eta\\,.\n\t\\label{tilts2}\n\\end{equation}\nThe above result is in-line with the data from the Planck satellite \\cite{Mukherjee:2015mma,Axelsson:2013mva}. \n\nIn Fig.~\\ref{fig:fig1} we depict the power asymmetry amplitude of the scalar power spectra for $n_s = 0.963$. This plot remains the same for any single-field slow-roll inflation.\n\\begin{figure}[ht]\n\t\\centering\n\t\\includegraphics[width=0.8\\linewidth]{Fig1}\n\t\\caption{In the above plot we depict the amplitude of CMB dipolar modulation or hemispherical asymmetry of scalar power spectra obtained from \\eqref{Ako}. In the plot we choose the pivot scale $k_\\ast=a_\\ast H_\\ast=0.05 {\\rm Mpc}^{-1}$ and fix $n_s=0.963$. The blue dot with error bar in the plot corresponds to the observational constraint $\\vert A \\vert = 0.066\\pm 0.021$ on HPA at large angular scales or at low-$\\ell-2-64$ or $k\\lesssim 10^{-1}k_\\ast$. We can notice from the plot that the HPA ceases to exist for the small angular scales or high-$\\ell$ which is compatible with the Planck satellite observations \\cite{Planck:2013lks,Akrami:2014eta,Aiola:2015rqa}. }\n\t\\label{fig:fig1}\n\\end{figure}\nSimilar to the scalar perturbations we propose the tensor perturbations to be direct sum of two kinds of fluctuations related by our definition of $\\mathcal{P}\\mathcal{T}$ transformation. \n\\begin{equation}\n\t\\hat{u}_{ij}\\left( \\tau,\\, \\textbf{x} \\right) = \\frac{1}{\\sqrt{2}}\t\\hat{u}_{\\rm I,\\, ij}\\left( \\tau,\\, \\textbf{x} \\right) \\oplus \\frac{1}{\\sqrt{2}} \\hat{u}_{\\rm II,\\, ij}\\left( -\\tau,\\, -\\textbf{x} \\right)\\,. \n\\end{equation}\nSimilar to the scalar power spectra we obtain two tensor power spectra which describe two point tensor correlations in the direction $\\hat{\\textbf{n}}$ and $-\\hat{\\textbf{n}}$ respectively. The two power spectra of tensor correlations are computed as \n\\begin{equation}\n\t\\begin{aligned}\n\t\tP_{h1} & = \\frac{k^3}{2\\pi^2}\\frac{4}{a^2} \\vert u_{\\rm I\\,k}\\vert^2\\Bigg\\vert_{\\tau = -\\frac{1}{aH}} \\\\ \n\t\t& \\approx \\frac{2H^2}{\\pi^2} \\left( 1+\\left( \\frac{k}{aH} \\right)^2 \\right) \\\\ &\\quad + \\frac{H^2}{\\pi}\\epsilon\\left(\\frac{k}{aH}\\right)^3 H_{3\/2}^{(1)} \\left( \\frac{k}{aH} \\right) \\frac{\\partial H^{(1)}_{\\nu_t}\\left( \\frac{k}{aH} \\right)}{\\partial \\nu_t}\\Big\\vert_{\\nu_t=3\/2}\\, \\\\\n\t\tP_{h2} & = \\frac{k^3}{2\\pi^2}\\frac{4}{a^2} \\vert u_{\\rm II,\\,k}\\vert^2\\Bigg\\vert_{\\tau_\\ast = \\frac{1}{aH}} \\\\ \n\t\t& \\approx \\frac{2H^2}{\\pi^2} \\left( 1+\\left( \\frac{k}{aH} \\right)^2 \\right)\\\\ &\\quad - \\frac{H^2}{\\pi} \\epsilon \\left(\\frac{k}{aH}\\right)^3 H_{3\/2}^{(1)} \\left( \\frac{k}{aH} \\right) \\frac{\\partial H^{(1)}_{\\nu_t}\\left( \\frac{k}{aH} \\right)}{\\partial \\nu_t}\\Big\\vert_{\\nu_t=3\/2}\\,. \n\t\\end{aligned}\n\t\\label{pwt12}\n\\end{equation}\nThe hemispherical power asymmetry of the tensor-power spectrum can be defined as \n\\begin{equation}\n\tT(k) = \\frac{P_{h1}-P_{h2}}{4P_h}\n\t\\label{Tko}\n\\end{equation}\nTo quantify the above amplitude $T(k)$ we choose Starobinsky or Higgs inflationary potential in Einstein frame \n\\begin{equation}\n\tV(\\phi) = \\frac{\\lambda}{4} \\left( 1- e^{-\\sqrt{\\frac{2}{3}}\\phi} \\right)^2\\,.\n\t\\label{staro-in}\n\\end{equation}\nwhich gives the slow-roll parameters in terms of e-folds as \\cite{Kehagias:2013mya}\n\\begin{equation}\n\t\\epsilon = \\frac{3}{4N^2},\\quad \\eta= \\frac{2}{N}\n\\end{equation}\nIn Fig.~\\ref{fig:fig2} we depict HPA amplitude of the tensor power spectra for the case of Starobinsky potential \\eqref{staro-in}.\n\\begin{figure}[ht]\n\t\\centering\n\t\\includegraphics[width=0.8\\linewidth]{Fig2}\n\t\\caption{In the above plot we depict the amplitude of dipolar modulation for primordial gravitational waves (PGWs) or HPA of tensor power spectra obtained from \\eqref{Tko}. The above plot is for the case of Starobinsky or Higgs inflation \\eqref{staro-in} for $N=55$ number of e-folds corresponding to the pivot scale $k_\\ast =0.05 {\\rm Mpc}^{-1}$.}\n\t\\label{fig:fig2}\n\\end{figure}\nSimilar the case of the tilt of scalar power spectra \\eqref{tilts2} we obtain the tilt of the tensor power spectra almost the same in the two opposite directions of the sky\n\\begin{equation}\n\t\\frac{d\\ln P_{h 1}}{d\\ln k}\\approx \t\\frac{d\\ln P_{h 2}}{d\\ln k} \\approx n_t \\approx -2\\epsilon\\,.\n\t\\label{tilts3}\n\\end{equation}\nFrom Fig.~\\ref{fig:fig1} and Fig.~\\ref{fig:fig2} we can notice that the HPA decreases as we increase the wave-number. This is a physically expected behavior because the short wavelength modes do not see the curvature of spacetime therefore the asymmetry should decrease with increasing $k$. In other words we also see it as how $(\\mathcal{C})\\mathcal{P}\\mathcal{T}$ symmetry is broken in a curved spacetime as a function of length scales. \n\n\n\\section{Conclusions}\n\nThe success of inflationary cosmology is its mapping of quantum fluctuations to the temperature fluctuations in the CMB. The HPA seems to pose a significant challenge to the standard understanding of inflationary quantum fluctuations because there is an asymmetry in the temperature correlations in the northern and southern hemispheres of the CMB sky. In this paper, we investigated whether the power asymmetry in the CMB can be originated from re-formulating our understanding of the quantization of inflationary fluctuations. Inspired from several open questions in the context of QFT in curved spacetime and quantum cosmology we proposed that the inflationary quantum fluctuations are generated in pairs which are related by discrete transformations of spacetime. Our approach is inspired from the fact that QFT in Minkowski spacetime is built on discrete symmetries such as $\\mathcal{C}\\mathcal{P}\\mathcal{T}$. In the case of curved spacetime i.e., \nwhen it involves gravity one must be careful in understanding spacetime reflection symmetries because gravity introduces dynamics. In our framework, we divide the notions of time into classical and quantum mechanical aspects. In the context of inflationary cosmology the standard procedure is to find a background geometry and quantize gravitational and matter degrees of freedom respecting the classical arrow of time \\cite{Martin:2004um}. However, we develop a new framework of quantization in canonical single field inflation where at the quantum level we allow the existence of quantum states that propagate both forward and backward in time based on the $\\mathcal{P}\\mathcal{T}$ transformations (since we only deal with chargeless fields we omit $\\mathcal{C}$ transformation) of the spacetime. Both of these quantum states in our framework co-exist within the co-moving horizon and they exit the horizon in the two opposite directions. We explicitly showed that these two type of quantum states even though created non-locally still do not violate any causality and locality principles in our formulation of QFT.\nWe showed that these pairs of quantum fluctuations behave exactly the same in the case of exact dS spacetime, but they are slightly different in the case of inflationary spacetime, which can be interpreted as the spontaneous breaking of $(\\mathcal{C})\\mathcal{P}\\mathcal{T}$ symmetry. This slight deviation seems to produce the power asymmetry hinted in the CMB observations. Our quantization scheme also applied to the case of inflationary tensor power spectrum and we predict a power asymmetry there as well. Notably, the HPA for scalar and tensor power spectrum is significant only for large angular scales or small wave numbers and it decreases for small angular scales or large wave numbers. This is expected since the spontaneous breaking of $(\\mathcal{C})\\mathcal{P}\\mathcal{T}$ symmetry is significant for small $k$, which are more affected by the background curved spacetime, compared to the large $k$.\nWe quantified all our predictions which serves as an observational test of our formalism. If the future observations \\cite{Li:2019bsg,Zhai:2017ibd} confirm these results we will learn significantly about the nature of inflationary quantum fluctuations as well as QFT in curved spacetime. \n \n\n\n\n\n\\begin{acknowledgments}\nKSK acknowledges the support from JSPS and KAKENHI Grant-in-Aid for Scientific Research No. JP20F20320, and thank Mainz U. for hospitality where part of the work has been carried out. J. Marto is supported by the grant UIDB\/MAT\/00212\/2020. We would like to thank Chris Ripken for the useful discussions and suggestions on quantization procedure and for the initial collaboration on the project. We thank Gerard 't Hooft for his inspiring talks and discussions related to QFT in curved spacetime. We thank Yashar Akrami, Norma G. Sanchez, Masahide Yamaguchi, Alexei A. Starobinsky, Luca Buoninfante, Francesco Di Fillippo, Yasha Neiman, Paolo Gondolo, Martin Reuter, Eiichiro Komatsu, Paulo V. Moniz, Dhiraz Kumar Hazra and L. Sriramkumar for very useful discussions.\n\\end{acknowledgments}\n\n\n\n\n\n\\bibliographystyle{apsrev4-2}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\\label{sec:intro}\n\nThis note reports the Tevatron average top-quark mass obtained by\ncombining the most precise published and preliminary measurements of\nthe top-quark mass. It is an update of the combination presented in\nRef.~\\cite{TeVTopComboPRD}, where further details can be found.\nThe ATLAS and CMS collaborations have also performed a combination of their most recent\ntop quark mass measurements~\\cite{lhccombi}.\n\nThe CDF and D\\O\\ collaborations have performed several direct experimental measurements of the\ntop-quark mass (\\ensuremath{M_{\\mathrm{t}}}) using data collected at the Tevatron\nproton-antiproton collider located at the Fermi National Accelerator\nLaboratory. These pioneering measurements were first based on approximately\n$0.1~\\ensuremath{\\mathrm{fb}^{-1}}$ of \\hbox{Run\\,I}\\ data\n~\\cite{Mtop1-CDF-di-l-PRLa}-\\cite{Mtop1-D0-allh-PRL}\ncollected from 1992 to 1996,\nand included results from the decay channels \\ensuremath{\\ttbar\\ra\\WbWb\\ra\\had}\\ (alljets),\n\\ensuremath{\\ttbar\\ra\\WbWb\\ra\\ljt}\\ ($\\ell$+jets), and \n\\ensuremath{\\ttbar\\ra\\WbWb\\ra\\dil}\\ ($\\ell\\ell$), where $\\ell=e$ or $\\mu$. \nDecays with $\\tau \\to e, \\mu$ are included in the direct $W \\to e$ and\n$W \\to \\mu$ channels. \nIn \\hbox{Run\\,II}\\ (2001--2011), many top mass measurements have been performed, and\nthose considered here are the most\nrecent results in these channels, using up to $8.7~\\ensuremath{\\mathrm{fb}^{-1}}$ of data for CDF\n(corresponding to the full CDF Run II dataset)\n~\\cite{\nMtop2-CDF-MEt-new,\nMtop2-CDF-l+jt-pub-new,\nMtop2-CDF-di-l-pub,\nMtop2-CDF-allh-pub},\nand up to $5.4~\\ensuremath{\\mathrm{fb}^{-1}}$ of data for D\\O\\\n~\\cite{ \nMtop2-D0-l+ja-final,\nMtop2-D0-l+jt-new2,\nMtop2-D0-di-l}.\nThe CDF analysis based upon charged particle tracking for exploiting\nthe transverse decay length \nof $b$-tagged jets ($L_{XY}$) and the transverse momentum of\nelectrons and muons from $W$ boson decays ($\\pte$) uses\na data set corresponding to a luminosity of\n1.9~$\\ensuremath{\\mathrm{fb}^{-1}}$~\\cite{Mtop2-CDF-trk}, and\nthere are no plans to update this analysis.\nThe D\\O\\ \\hbox{Run\\,II}\\ measurements presented in this note include the most\nrecent Run II measurement in \nthe $\\ell\\ell$~\\cite{Mtop2-D0-di-l} channel using 5.4~fb$^{-1}$ of data \nand in the $\\ell$+jets channel \\cite{Mtop2-D0-l+jt-new2} \nwith 3.6~fb$^{-1}$ of data. Both results are now published.\nSince the combination performed in 2011~\\cite{Mtop-tevewwgSum11}, a new\nfinal state signature was introduced by CDF that requires events to\npossess missing transverse energy ($\\met$) and jets, but no identified\nlepton (``MEt'')~\\cite{Mtop2-CDF-MEt-new,Mtop2-CDF-MEt-published}. \nThis sample is statistically independent from the previous three CDF\nchannels.\n\nWith respect to the July 2011 combination\\,\\cite{Mtop-tevewwgSum11} and the published version of the \ncombination~\\cite{TeVTopComboPRD}, the \\hbox{Run\\,II}\\ CDF measurement in the \n$\\ell$+jets channel has been updated using 8.7~fb$^{-1}$ \nof data, an improved analysis technique, and improved jet energy resolution \\cite{Mtop2-CDF-l+jt-pub-new}. \nThe CDF measurement in the MEt channel was updated to use the full\n\\hbox{Run\\,II}\\ data set for CDF of 8.7~fb$^{-1}$ of data as \nwell~\\cite{Mtop2-CDF-MEt-new}.\nThe now published \\hbox{Run\\,II}\\ CDF measurements in the $\\ell\\ell$ channel\\,\\cite{Mtop2-CDF-di-l-pub} and alljets \nchannel\\,\\cite{Mtop2-CDF-allh-pub} are unchanged. The measurement \nbased on charged particle tracking~\\cite{Mtop2-CDF-trk} was\nincorporated as described in the past \ncombinations~\\cite{Mtop-tevewwgSum11}. \nFrom the corresponding analysis only\nthe measurement of the top quark mass using the mean decay length $\\Lxy$ of\n$B$ hadrons in $b$-tagged lepton+jets events has been used. It\nis independent of energy information in the calorimeter, and its main\nsource of systematic uncertainty is uncorrelated with the dominant\nones from the jet energy scale calibration in other measurements. This\nmeasurement of $m_t$ is essentially uncorrelated with the higher\nprecision CDF result from the lepton+jets channel. The overlap between\nthe data samples used for the decay-length method and the lepton+jets\nsample has therefore no effect. \n\nThe Tevatron average top-quark mass is obtained by combining five published\n\\hbox{Run\\,I}\\ measurements~\\cite{Mtop1-CDF-di-l-PRLb, Mtop1-CDF-di-l-PRLb-E,\n Mtop1-D0-di-l-PRD, Mtop1-CDF-l+jt-PRD, Mtop1-D0-l+jt-new1,\n Mtop1-CDF-allh-PRL} with four published \\hbox{Run\\,II}\\ CDF \nresults~\\cite{Mtop2-CDF-l+jt-pub-new,Mtop2-CDF-di-l-pub,Mtop2-CDF-allh-pub, \nMtop2-CDF-trk}, one preliminary \\hbox{Run\\,II}\\ CDF\nresult~\\cite{Mtop2-CDF-MEt-new}, and two published \n\\hbox{Run\\,II}\\ D\\O\\ results~\\cite{Mtop2-D0-l+jt-new2,\nMtop2-D0-di-l}.\nThis combination \nsupersedes previous\ncombinations~\\cite{Mtop-tevewwgSum11,Mtop1-tevewwg04,Mtop-tevewwgSum05,\n Mtop-tevewwgWin06,Mtop-tevewwgSum06, Mtop-tevewwgWin07, Mtop-tevewwgWin08, \n Mtop-tevewwgSum08, Mtop-tevewwgWin09,Mtop-tevewwgSum10}. \n\nThe definition and evaluation of the systematic uncertainties and the understanding of the\ncorrelations among channels, experiments, and Tevatron runs is the outcome of many years of \njoint work between the CDF and D\\O\\ collaborations and is described in detail\nelsewhere~\\cite{TeVTopComboPRD}.\n\nThe input measurements and uncertainty categories used in the combination are \ndetailed in Sections~\\ref{sec:inputs} and~\\ref{sec:uncertainty}, respectively. \nThe correlations assumed in the combination are discussed in \nSection~\\ref{sec:corltns} and the resulting Tevatron average top-quark mass \nis given in Section~\\ref{sec:results}. A summary is presented\nin Section~\\ref{sec:summary}.\n \n\\section{Input Measurements}\n\\label{sec:inputs}\n\nTwelve measurements of \\ensuremath{M_{\\mathrm{t}}}\\ used in this combination are shown in Table~\\ref{tab:inputs}.\nThe \\hbox{Run\\,I}\\ measurements all have relatively large statistical\nuncertainties and their systematic uncertainties are dominated by the\ntotal jet energy scale (JES) uncertainty. In \\hbox{Run\\,II}\\, both CDF and\nD\\O\\ take advantage of the larger \\ensuremath{t\\overline{t}}\\ samples available and employ\nnew analysis techniques to reduce both of these uncertainties. In\nparticular, the \\hbox{Run\\,II}\\ D\\O\\ analysis in the $\\ell$+jets channel and the \n\\hbox{Run\\,II}\\ CDF analyses in the $\\ell$+jets, alljets, and MEt channels \nconstrain the response of light-quark jets using the kinematic information from $W\\ra\nqq^{\\prime}$ decays (so-called {\\em in situ}\ncalibration)~\\cite{Mtop1-CDF-l+jt-PRD,Abazov:2006bd}. Residual JES uncertainties associated with\n$p_{T}$ and $\\eta$ dependencies as well as uncertainties specific to\nthe response of $b$ jets are treated separately. The\n\\hbox{Run\\,II}\\ D\\O\\ $\\ell\\ell$ measurement uses the JES determined in the\n$\\ell$+jets channel by {\\em in situ} calibration~\\cite{Mtop2-D0-di-l}.\n\n\\vspace*{0.10in}\n\n\\begin{table}[t]\n\\caption[Input measurements]{Summary of the measurements used to determine the\n Tevatron average \\ensuremath{M_{\\mathrm{t}}}. Integrated luminosity ($\\int \\mathcal{L}\\;dt$) has units of\n \\ensuremath{\\mathrm{fb}^{-1}}, and all other numbers are in $\\ensuremath{\\mathrm{ Ge\\kern -0.1em V }\\kern -0.2em \/c^2 }$. The uncertainty categories and \n their correlations are described in Section~\\ref{sec:uncertainty}. The total systematic uncertainty \n and the total uncertainty are obtained by adding the relevant contributions \n in quadrature. ``n\/a'' stands for ``not applicable'', ``n\/e'' for ``not evaluated''.}\n\\label{tab:inputs}\n\\begin{center}\n\n\\renewcommand{\\arraystretch}{1.30}\n{\\tiny\n\\begin{tabular}{l|ccc|cc|cccc|cc|c} \n\\hline \\hline\n & \\multicolumn{5}{c|}{{\\hbox{Run\\,I}} published} \n & \\multicolumn{6}{c|}{{\\hbox{Run\\,II}} published} \n & \\multicolumn{1}{c}{{\\hbox{Run\\,II}} prel.} \\\\ \n \n & \\multicolumn{3}{c|}{ CDF } \n & \\multicolumn{2}{c}{ D\\O\\ }\n & \\multicolumn{4}{|c|}{ CDF }\n & \\multicolumn{2}{c|}{ D\\O\\ }\n & \\multicolumn{1}{c}{ CDF }\n \\\\\n\n & $\\ell$+jets & $\\ell\\ell$ & alljets & $\\ell$+jets & $\\ell\\ell$ & $\\ell$+jets & $\\ell\\ell$ & alljets & Lxy & $\\ell$+jets & $\\ell\\ell$ & MEt \\\\\n\\hline \n$\\int \\mathcal{L}\\;dt$& 0.1 & 0.1 & 0.1 & 0.1 & 0.1 & 8.7 & 5.6 & 5.8 & 1.9 & 3.6 & 5.3 & 8.7 \\\\\n\\hline \nResult & 176.1 & 167.4 &186.0 & 180.1 & 168.4 & 172.85 & 170.28 & 172.47 & 166.90 & 174.94 & 174.00 & 173.95\\\\\n\\hline \n\\shortstack{ {\\em In situ} light-jet cali- \\\\\nbration (iJES)} & n\/a & n\/a & n\/a & n\/a & n\/a & 0.49 & n\/a & 0.95 & n\/a & 0.53 & 0.55 & 1.05 \\\\\n\\shortstack{ Response to $b$\/$q$\/$g$ \\\\\njets (aJES)} & n\/a & n\/a & n\/a & 0.0 & 0.0 & 0.09 & 0.14 & 0.03 & n\/a & 0.0 & 0.40 & 0.10 \\\\\n\\shortstack{ Model for $b$ jets \\\\\n(bJES)} & 0.6 & 0.8 & 0.6 & 0.7 & 0.7 & 0.16 & 0.33 & 0.15 & n\/a & 0.07 & 0.20 & 0.17 \\\\\n \\shortstack{ Out-of-cone correction \\\\\n (cJES)} & 2.7 & 2.6 & 3.0 & 2.0 & 2.0 & 0.21 & 2.13 & 0.24 & 0.36 & n\/a & n\/a & 0.18 \\\\\n\\shortstack{ Light-jet response (2) \\\\ \n (dJES)} & 0.7 & 0.6 & 0.3 & 2.5 & 1.1 & 0.07 & 0.58 & 0.04 & 0.06 & 0.63 & 0.56 & 0.04 \\\\\n\\shortstack{ Light-jet response (1) \\\\\n (rJES)} & 3.4 & 2.7 & 4.0 & n\/a & n\/a & 0.48 & 2.01 & 0.38 & 0.24 & n\/a & n\/a & 0.40 \\\\\n\\shortstack{ Lepton modeling \\\\\n(LepPt)} & n\/e & n\/e & n\/e & n\/e & n\/e & 0.03 & 0.27 & n\/a & n\/a & 0.17 & 0.35 & n\/a \\\\\n\\shortstack{ Signal modeling \\\\\n (Signal)} & 2.6 & 2.9 & 2.0 & 1.1 & 1.8 & 0.61 & 0.73 & 0.62 & 0.90 & 0.77 & 0.86 & 0.64\\\\\n \\shortstack{ Jet modeling \\\\\n (DetMod)} & 0.0 & 0.0 & 0.0 & 0.0 & 0.0 & 0.0 & 0.0 & 0.0 & 0.0 & 0.36 & 0.50 & 0.0 \\\\\n\\shortstack{ Offset \\hspace{0.5cm} \\\\\n (UN\/MI)} & n\/a & n\/a & n\/a & 1.3 & 1.3 & n\/a & n\/a & n\/a & n\/a & n\/a & n\/a & n\/a \\\\\n\\shortstack{ Background from \\\\\ntheory (BGMC)} & 1.3 & 0.3 & 1.7 & 1.0 & 1.1 & 0.12 & 0.24 & 0.0 & 0.80 & 0.18 & 0.0 & 0.0 \\\\\n \\shortstack{ Background based on\\\\\ndata (BGData)} & 0.0 & 0.0 & 0.0 & 0.0 & 0.0 & 0.16 & 0.14 & 0.56 & 0.20 & 0.23 & 0.20 & 0.12 \\\\\n \\shortstack{ Calibration method\\\\\n(Method)} & 0.0 & 0.7 & 0.6 & 0.6 & 1.1 & 0.00 & 0.12 & 0.38& 2.50 & 0.16 & 0.51 & 0.31 \\\\\n \\shortstack{ Multiple interactions\\\\\n model (MHI)} & n\/e & n\/e & n\/e & n\/e & n\/e & 0.07 & 0.23 & 0.08& 0.0 & 0.05 & 0.0 & 0.18 \\\\\n\\hline \n \\shortstack{ Systematic uncertainty\\\\\n (Syst)} & 5.3 & 4.9 & 5.7 & 3.9 & 3.6 & 0.98 & 3.09 & 1.49 & 2.90 & 1.24 & 1.44 & 1.35 \\\\\n\\shortstack{ Statistical uncertainty\\\\\n (Stat)} & 5.1 & 10.3 &10.0 & 3.6 & 12.3 & 0.52 & 1.95 & 1.43 & 9.00 & 0.83 & 2.36 & 1.26 \\\\\n\\hline \nTotal uncertainty & 7.3 & 11.4 &11.5 & 5.3 & 12.8 & 1.11 & 3.79 & 2.06 & 9.46 & 1.50 & 2.76 & 1.85\\\\ \n\\hline\n\\hline\n\\end{tabular}\n}\n\\end{center}\n\\end{table}\n\n\nThe D\\O\\ Run~II $\\ell$+jets analysis uses the JES determined from the\nexternal calibration derived from $\\gamma$+jets events as an\nadditional Gaussian constraint to the {\\em in situ} calibration. Therefore,\nthe total resulting JES uncertainty is split into one part obtained from the \n{\\em in situ} calibration and another part determined from the external calibration.\nTo do this, the measurement without external\nJES constraint has been combined iteratively with a pseudo-measurement\nusing the method of Refs.~\\cite{Lyons:1988, Valassi:2003} \nthat uses only the external calibration in a way that the combination gives the total JES uncertainty. \nThe splitting obtained in this way is used to assess both the statistical part of the JES uncertainty and the part of \nthe JES uncertainty due to the external calibration constraint~\\cite{Mtop2-D0-comb}.\n\nThe $\\Lxy$ technique developed by CDF \nuses the decay length of $B$ mesons from $b$-tagged jets.\nWhile the statistical sensitivity of this analysis is not as good as\nthat of the more\ntraditional methods, this technique has the advantage that it is almost entirely independent of\nJES uncertainties since it\nuses primarily tracking information.\n\\vspace*{0.10in}\n\nThe D\\O\\ \\hbox{Run\\,II}\\ $\\ell$+jets result is a combination of the \npublished Run~IIa (2002--2005) measurement ~\\cite{Mtop2-D0-l+ja-final} with 1~fb$^{-1}$ \nof data and the result obtained with 2.6~fb$^{-1}$ of data from Run~IIb (2006--2007)~\\cite{Mtop2-D0-l+jt-new2}.\nThis analysis includes an additional particle response correction on top of the\nstandard {\\it in situ} calibration.\nThe D\\O\\ \\hbox{Run\\,II}\\ $\\ell\\ell$ result is based on a neutrino weighting technique using 5.4~fb$^{-1}$\nof \\hbox{Run\\,II}\\ data~\\cite{Mtop2-D0-di-l}. \n\\vspace*{0.10in}\n\nTable~\\ref{tab:inputs} lists the individual uncertainties of each result,\nsubdivided into the categories described in the next Section. The\ncorrelations between the inputs are described in\nSection~\\ref{sec:corltns}.\n\n\n\n\\section{Uncertainty Categories}\n\\label{sec:uncertainty}\n\nWe employ uncertainty categories similar to what was used for the previous Tevatron\naverage~\\cite{TeVTopComboPRD,Mtop-tevewwgSum11}, with small\nmodifications to better account for their correlations.\nThey are divided such that sources of systematic uncertainty that share the same or similar origin are \ncombined as explained in Ref.~\\cite{TeVTopComboPRD}. \nFor example, the {\\it Signal modeling} ({\\it Signal}) category\ndiscussed below includes the uncertainties from different systematic\nsources that are correlated due to their origin in the modeling of the simulated signal samples.\n\n Some systematic uncertainties have been separated into multiple\ncategories to accommodate specific types of correlations.\nFor example, the jet energy scale (JES) uncertainty is subdivided\ninto six components to more accurately accommodate our\nbest understanding of the relevant correlations between input measurements. \n\nFor this note we use the new systematic naming scheme described in Ref.~\\cite{TeVTopComboPRD}. In parentheses, the \nold names of the systematic uncertainties are provided. There is a one-to-one matching between the new and old systematic \ndefinitions of categories.\n\n\\vspace*{0.10in}\n\n\\begin{description}\n \\item[Statistical uncertainty (Statistics):] The statistical uncertainty associated with the\n \\ensuremath{M_{\\mathrm{t}}}\\ determination.\n \\item[{\\em In situ} light-jet calibration (iJES):] That part of the\n JES uncertainty that originates from\n {\\em in situ} calibration procedures and is uncorrelated among the\n measurements. In the combination reported here, it corresponds to\n the statistical uncertainty associated with the JES determination\n using the $W\\ra qq^{\\prime}$ invariant mass in the CDF \\hbox{Run\\,II}\\\n $\\ell$+jets, alljets, and MEt measurements and the D\\O\\ Run~II\n $\\ell\\ell$ and $\\ell$+jets\n measurements. \n For the D\\O\\ Run~II $\\ell$+jets measurement, it also includes the uncertainty\n coming from the MC\/data difference in jet response that is uncorrelated\n with the other D\\O\\ Run~II measurements. \n Residual JES uncertainties arising from effects\n not considered in the {\\em in situ} calibration are included in other\n categories. \n \\item[Response to \\boldmath{$b\/q\/g$} jets (aJES):] That part of the JES\n uncertainty that originates from \n average differences in detector electromagnetic over hadronic ($e\/h$)\n response for hadrons produced in the fragmentation of $b$-jets and light-quark \n jets. \n \\item[Model for \\boldmath{$b$} jets (bJES):] That part of the JES uncertainty that originates from\n uncertainties specific to the modeling of $b$ jets and that is correlated\n across all measurements. For both CDF and D\\O\\ this includes uncertainties \n arising from \n variations in the semileptonic branching fractions, $b$-fragmentation \n modeling, and differences in the color flow between $b$-quark jets and light-quark\n jets. These were determined from \\hbox{Run\\,II}\\ studies but back-propagated\n to the \\hbox{Run\\,I}\\ measurements, whose {\\it Light-jet response (1)} uncertainties ({\\it rJES}, see below) were \n then corrected to keep the total JES uncertainty constant.\n \\item[Out-of-cone correction (cJES):] That part of the JES uncertainty that originates from\n modeling uncertainties correlated across all measurements. \n It specifically includes the modeling uncertainties associated with light-quark \n fragmentation and out-of-cone corrections. For D\\O\\ \\hbox{Run\\,II}\\ measurements,\n it is included in the {\\it Light-jet response (2) (dJES)} category.\n \\item[Light-jet response (1) (rJES):] The remaining part of the JES\n uncertainty that covers the absolute calibration for CDF's \\hbox{Run\\,I}\\\n and \\hbox{Run\\,II}\\ measurements. It also includes small contributions\n from the uncertainties associated with modeling multiple\n interactions within a single bunch crossing and corrections for\n the underlying event. \n \n \n \n \n \n \n \\item[Light-jet response (2) (dJES):] That part of the JES\n uncertainty that includes D\\O's \\hbox{Run\\,I}\\ and \\hbox{Run\\,II}\\ calibrations of\n absolute response (energy dependent), the relative response\n ($\\eta$-dependent), and the out-of-cone showering correction that \n is a detector effect. This uncertainty term for CDF includes only\n the small relative response calibration ($\\eta$-dependent) for\n \\hbox{Run\\,I}\\ and \\hbox{Run\\,II}. \n \n \n \n \n \n \n \n \n \n \n \n \\item[Lepton modeling (LepPt):] The systematic uncertainty arising from uncertainties\n in the scale of lepton transverse momentum measurements. It was not\n considered as a source of systematic uncertainty in the \\hbox{Run\\,I}\\\n measurements. \n \\item[Signal modeling (Signal):] The systematic uncertainty arising from uncertainties\n in \\ensuremath{t\\overline{t}}\\ modeling that is correlated across all\n measurements. This includes uncertainties from variations of the amount of initial and \n final state radiation and from the choice of parton density function used\n to generate the \\ensuremath{t\\overline{t}}\\ Monte Carlo samples\n that calibrate each method. For D\\O, it also includes the uncertainty \n from higher-order corrections evaluated from a comparison of \\ensuremath{t\\overline{t}}\\ samples generated by {\\textsc MC@NLO} ~\\cite{MCNLO} and \n {\\textsc ALPGEN}~\\cite{ALPGEN}, both interfaced to {\\textsc HERWIG}~\\cite{HERWIG5,HERWIG6} for the simulation of parton showers and hadronization.\n In this combination, the systematic uncertainty arising from a variation of the \n phenomenological description of color reconnection (CR) between final state particles \\cite{CR,Skands:2009zm} \n is included in the {\\it Signal modeling} category.\n The CR uncertainty is obtained by taking the difference between the {\\textsc PYTHIA}\\,6.4 tune ``Apro\" and the {\\textsc PYTHIA}\\,6.4 tune \n``ACRpro\" that differ only in the CR model. \nThis uncertainty was not evaluated in Run~I since the Monte Carlo\ngenerators available at that time did not allow for variations of the CR model.\nThese measurements therefore do not include this source of systematic uncertainty. Finally, the systematic \nuncertainty associated with variations of the MC generator used to calibrate the mass extraction method is added. \nIt includes variations observed when \n substituting {\\textsc PYTHIA}\n \\cite{PYTHIA4,PYTHIA5,PYTHIA6} \n (\\hbox{Run\\,I}\\ and \\hbox{Run\\,II}) \n or {\\textsc ISAJET}~\\cite{ISAJET} (\\hbox{Run\\,I}) for {\\textsc HERWIG}~\\cite{HERWIG5,HERWIG6} when \n modeling the \\ensuremath{t\\overline{t}}\\ signal. \n \\item[Jet modeling (DetMod):] The systematic uncertainty arising from uncertainties \nin the modeling of jet interactions in the detector in the MC\nsimulation. For D\\O\\, this includes uncertainties from jet resolution\nand identification. \nApplying jet algorithms to MC events, CDF finds that the resulting\nefficiencies and resolutions closely match those in data. The small\ndifferences propagated to $\\ensuremath{M_{\\mathrm{t}}}$ lead to a negligible uncertainty of\n0.005~GeV, which is then ignored. \n\n \\item[Background based on data (BGData):] This includes \n uncertainties associated with the modeling using data of the QCD\n multijet background in the alljets,\n MEt, and $\\ell$+jets channels and\n the Drell-Yan background in the $\\ell\\ell$ channel. \n This part is uncorrelated between experiments.\n\n \\item[Background from theory (BGMC):] \nThis systematic uncertainty on the background originating from theory\n(MC) takes into account the \n uncertainty in modeling the background sources. It is correlated between\n all measurements in the same channel, and includes uncertainties on the background composition, normalization, and shape of different components, e.g., the \nuncertainties from the modeling of the $W$+jets background in the $\\ell$+jets channel \nassociated with variations of the factorization scale used to simulate $W$+jets events. \n\n \\item[Calibration method (Method):] The systematic uncertainty arising from any source specific\n to a particular fit method, including the finite Monte Carlo statistics \n available to calibrate each method. \n \\item[Offset (UN\/MI):] This uncertainty is specific to D\\O\\ and includes the uncertainty\n arising from uranium noise in the D\\O\\ calorimeter and from the\n multiple interaction corrections to the JES. For D\\O\\ \\hbox{Run\\,I}\\ these\n uncertainties were sizable, while for \\hbox{Run\\,II}, owing to the shorter\n calorimeter electronics integration time and {\\em in situ} JES calibration, these uncertainties\n are negligible.\n \\item[Multiple interactions model (MHI):] The systematic uncertainty arising from a mismodeling of \n the distribution of the number of collisions per Tevatron bunch crossing owing to the \n steady increase in the collider instantaneous luminosity during data-taking. \n This uncertainty has been separated from other sources to account for the fact that \n it is uncorrelated between experiments.\n\n\\end{description}\nThese categories represent the current preliminary understanding of the\nvarious sources of uncertainty and their correlations. We expect these to \nevolve as we continue to probe each method's sensitivity to the various \nsystematic sources with improving precision.\n\n\\section{Correlations}\n\\label{sec:corltns}\n\nThe following correlations are used for the combination:\n\\begin{itemize}\n \\item The uncertainties in the {\\it Statistical uncertainty (Stat)} and\n {\\it Calibration method (Method)}\n categories are taken to be uncorrelated among the measurements.\n \\item The uncertainties in the {\\it In situ light-jet \n calibration (iJES)}\n category are taken to be uncorrelated among the measurements\n except for D0's $\\ell\\ell$ and $\\ell$+jets measurements, where\n this uncertainty is taken to be 100\\% correlated since the\n $\\ell\\ell$ measurement uses the JES calibration determined in\n $\\ell$+jets channel. \n \\item The uncertainties in the {\\it Response to $b$\/$q$\/$g$ jets (aJES)}, {\\it Light-jet response (2) (dJES)}, {\\it Lepton modeling (LepPt)},\n and {\\it Multiple interactions model (MHI)} categories are taken\n to be 100\\% correlated among all \\hbox{Run\\,I}\\ and all \\hbox{Run\\,II}\\ measurements \n within the same experiment, but uncorrelated between \\hbox{Run\\,I}\\ and \\hbox{Run\\,II}\\\n and uncorrelated between the experiments.\n \\item The uncertainties in the {\\it Light-jet response (1) (rJES)}, {\\it Jet modeling (DetMod)}, and {\\it Offset (UN\/MI)} categories are taken\n to be 100\\% correlated among all measurements within the same experiment \n but uncorrelated between the experiments.\n \\item The uncertainties in the {\\it Backgrounds estimated from theory (BGMC)} category are taken to be\n 100\\% correlated among all measurements in the same channel.\n \\item The uncertainties in the {\\it Backgrounds estimated from data (BGData)} category are taken to be\n 100\\% correlated among all measurements in the same channel and same run period, but uncorrelated between the experiments.\n \\item The uncertainties in the {\\it Model for $b$ jets (bJES)}, {\\it Out-of-cone correction (cJES)}, and {\\it Signal modeling (Signal)}\n categories are taken to be 100\\% correlated among all measurements.\n\\end{itemize}\nUsing the inputs from Table~\\ref{tab:inputs} and the correlations specified\nhere, the resulting matrix of total correlation coefficients is given in\nTable~\\ref{tab:coeff}.\n\n\\begin{table}[t]\n\\caption[Global correlations between input measurements]{The matrix of correlation coefficients used to determine the\n Tevatron average top-quark mass.}\n\\begin{center}\n\\renewcommand{\\arraystretch}{1.30}\n\\tiny\n\\begin{tabular}{l|ccc|cc|cccc|cc|c}\n\\hline \\hline\n & \\multicolumn{5}{c}{{\\hbox{Run\\,I}} published} \n & \\multicolumn{6}{|c|}{{\\hbox{Run\\,II}} published} \n & \\multicolumn{1}{c}{{\\hbox{Run\\,II}} preliminary} \\\\\n & \\multicolumn{3}{c|}{ CDF } \n & \\multicolumn{2}{c}{ D\\O\\ }\n & \\multicolumn{4}{|c|}{ CDF } \n\t & \\multicolumn{2}{c|}{ D\\O\\ }\n & \\multicolumn{1}{c}{ CDF } \n \\\\\n\n & $\\ell$+jets & $\\ell\\ell$ & alljets & $\\ell$+jets & $\\ell\\ell$ & $\\ell$+jets & $\\ell\\ell$ & alljets & $\\Lxy$ & $\\ell$+jets & $\\ell\\ell$ & MEt \\\\\n \n\\hline \nCDF-I $\\ell$+jets & 1.00 & 0.29 & 0.32 & 0.26 & 0.11 & 0.49 & 0.54 & 0.25 & 0.07 & 0.21 & 0.12 & 0.27 \\\\\nCDF-I $\\ell\\ell$ & 0.29 & 1.00 & 0.19 & 0.15 & 0.08 & 0.29 & 0.32 & 0.15 & 0.04 & 0.13 & 0.08 & 0.17 \\\\\nCDF-I alljets & 0.32 & 0.19 & 1.00 & 0.14 & 0.07 & 0.30 & 0.38 & 0.15 & 0.04 & 0.09 & 0.06 & 0.16 \\\\\nD\\O-I $\\ell$+jets & 0.26 & 0.15 & 0.14 & 1.00 & 0.16 & 0.22 & 0.27 & 0.12 & 0.05 & 0.14 & 0.07 & 0.12 \\\\\nD\\O-I $\\ell\\ell$ & 0.11 & 0.08 & 0.07 & 0.16 & 1.00 & 0.11 & 0.13 & 0.07 & 0.02 & 0.07 & 0.05 & 0.07 \\\\\nCDF-II $\\ell$+jets & 0.49 & 0.29 & 0.30 & 0.22 & 0.11 & 1.00 & 0.48 & 0.29 & 0.08 & 0.30 & 0.18 & 0.33 \\\\\nCDF-II $\\ell\\ell$ & 0.54 & 0.32 & 0.38 & 0.27 & 0.13 & 0.48 & 1.00 & 0.25 & 0.06 & 0.11 & 0.07 & 0.26 \\\\\nCDF-II alljets & 0.25 & 0.15 & 0.15 & 0.12 & 0.07 & 0.29 & 0.25 & 1.00 & 0.04 & 0.16 & 0.10 & 0.17 \\\\\nCDF-II $\\Lxy$ & 0.07 & 0.04 & 0.04 & 0.05 & 0.02 & 0.08 & 0.06 & 0.04 & 1.00 & 0.06 & 0.03 & 0.04 \\\\\nD\\O-II $\\ell$+jets & 0.21 & 0.13 & 0.09 & 0.14 & 0.07 & 0.30 & 0.11 & 0.16 & 0.06 & 1.00 & 0.39 & 0.18 \\\\\nD\\O-II $\\ell\\ell$ & 0.12 & 0.08 & 0.06 & 0.07 & 0.05 & 0.18 & 0.07 & 0.10 & 0.03 & 0.39 & 1.00 & 0.11 \\\\\nCDF-II MEt & 0.27 & 0.17 & 0.16 & 0.12 & 0.07 & 0.33 & 0.26 & 0.17 & 0.04 & 0.18 & 0.11 & 1.00 \\\\\n\n\n\\hline\n\\hline\n\\end{tabular}\n\\end{center}\n\\label{tab:coeff}\n\\end{table}\n\nThe measurements are combined using a program implementing two \nindependent methods: \na numerical $\\chi^2$ minimization and \nthe analytic best linear unbiased estimator (BLUE) method~\\cite{Lyons:1988, Valassi:2003}. \nThe two methods are mathematically equivalent.\nIt has been checked that they give identical results for\nthe combination. The BLUE method yields the decomposition of the uncertainty on the Tevatron $\\ensuremath{M_{\\mathrm{t}}}$ average in \nterms of the uncertainty categories specified for the input measurements~\\cite{Valassi:2003}.\n\n\\section{Results}\n\\label{sec:results}\n\nThe resultant combined value for the top-quark mass is\n\\begin{eqnarray}\n\\nonumber\n\\ensuremath{M_{\\mathrm{t}}}=\\gevcc{\\measStatSyst{173.20}{0.51}{0.71}}. \n\\end{eqnarray}\nAdding the statistical and systematic uncertainties\nin quadrature yields a total uncertainty of $\\gevcc{0.87}$, corresponding to a\nrelative precision of 0.50\\% on the top-quark mass.\nIt has a $\\chi^2$ of 8.5 for 11 degrees of freedom, corresponding to\na probability of 67\\%, indicating good agreement among all input\nmeasurements. The breakdown of the uncertainties is \nshown in Table\\,\\ref{tab:BLUEuncert}. The total statistical and systematic \nuncertainties are reduced relative to the \nSummer 2011 combination \\cite{Mtop-tevewwgSum11} and the published combination~\\cite{TeVTopComboPRD} \ndue to the increase of the CDF data samples in the $\\ell$+jets and MEt analyses and better \ntreatment of JES corrections in the $\\ell$+jets analysis. \n \nThe pull and weight for each of the inputs, as obtained from the\ncombination with the BLUE method, are listed in Table~\\ref{tab:stat}.\nThe input measurements and the resulting Tevatron average mass of the top \nquark are summarized in Fig.~\\ref{fig:summary}.\n\\vspace*{0.10in}\n\n\\begin{table}[tbh]\n\\caption{\\label{tab:BLUEuncert} \nSummary of the Tevatron combined average $\\ensuremath{M_{\\mathrm{t}}}$. The uncertainty categories are \ndescribed in the text. The total systematic uncertainty and the total\nuncertainty are obtained \nby adding the relevant contributions in quadrature.}\n\\begin{center}\n\\begin{tabular}{lc} \\hline \\hline\n & Tevatron combined values (GeV\/$c^2$) \\\\ \\hline\n $\\ensuremath{M_{\\mathrm{t}}}$ & 173.20 \\\\ \\hline\n {\\em In situ} light-jet calibration (iJES) & 0.36 \\\\\n Response to $b$\/$q$\/$g$ jets (aJES) & 0.09 \\\\\n Model for $b$ jets (bJES) & 0.11 \\\\ \n Out-of-cone correction (cJES) & 0.01 \\\\ \n Light-jet response (2) (dJES) & 0.15 \\\\ \n Light-jet response (1) (rJES) & 0.16 \\\\\n Lepton modeling (LepPt) & 0.05 \\\\\n Signal modeling (Signal) & 0.52 \\\\\n Jet modeling (DetMod) & 0.08 \\\\\n Offset (UN\/MI) & 0.00 \\\\\n Background from theory (BGMC) & 0.06\\\\\n Background based on data (BGData) & 0.13\\\\\n Calibration method (Method) & 0.06 \\\\\n Multiple interactions model (MHI) & 0.07 \\\\ \\hline\n Systematic uncertainty (syst) & 0.71 \\\\ \n Statistical uncertainty (stat) & 0.51 \\\\ \\hline\n Total uncertainty & 0.87 \\\\\n \\hline \\hline\n\\end{tabular}\n\\end{center}\n\\end{table}\n\nThe weights of some of the measurements are negative, which occurs if\nthe correlation between two measurements \nis larger than the ratio of their total uncertainties.\nIn these instances the less precise measurement \nwill acquire a negative weight. While a weight of zero means that a\nparticular input is effectively ignored in the combination, channels\nwith a negative weight affect the resulting $\\ensuremath{M_{\\mathrm{t}}}$ central value and help reduce the total\nuncertainty~\\cite{Lyons:1988}. \nTo visualize the weight each measurement carries in the combination, Fig.\\,\\ref{fig:Weights} shows the\nabsolute values of the weight of each measurement divided by the sum of the absolute values of the weights\nof all input measurements. Negative weights are represented by bins\nwith a different (grey) color. \nWe note, that due to correlations between the uncertainties the relative weights\nof the different input channels may be significantly different from\nwhat one could expect from the total accuracy of each measurement as represented by\nerror bars in Fig.~\\ref{fig:summary}.\n\n\\begin{figure}[p]\n\\begin{center}\n\\includegraphics[width=0.8\\textwidth]{figures\/TevMtWin13-2digit.pdf}\n\\end{center}\n\\caption[Summary plot for the Tevatron average top-quark mass]\n {Summary of the input measurements and resulting Tevatron average\n mass of the top quark.}\n\\label{fig:summary} \n\\end{figure}\n\n\\begin{figure}\n\\begin{center}\n\\includegraphics[width=1.0\\textwidth]{figures\/TevWeightPlotWin13-neg.pdf}\n\\end{center}\n\\caption{Relative weights of the input measurements in the\n combination. The relative weights have been obtained by dividing the absolute value of each measurement weight by the sum over all measurements of the absolute values\nof the weights. Negative weights are represented by their absolute\nvalue, but using a grey color.}\n\\label{fig:Weights} \n\\end{figure}\n\n\\begin{table}[t]\n\\caption[Pull and weight of each measurement]{The pull and weight for each of the\n inputs, as obtained from the combination with the BLUE method to\n determine the average top quark mass.}\n\\begin{center}\n\\renewcommand{\\arraystretch}{1.30}\n{\\tiny\n\\begin{tabular}{l|ccccc|cccccc|c}\n\\hline \n\\hline\n & \\multicolumn{5}{c}{{\\hbox{Run\\,I}} published} \n & \\multicolumn{6}{|c|}{{\\hbox{Run\\,II}} published} \n & \\multicolumn{1}{c}{{\\hbox{Run\\,II}} preliminary} \\\\\n & \\multicolumn{3}{c}{ CDF } \n & \\multicolumn{2}{c}{ D\\O\\ }\n & \\multicolumn{4}{|c}{ CDF } \n & \\multicolumn{2}{c|}{ D\\O\\ } \n & \\multicolumn{1}{c}{ CDF } \\\\\n & $\\ell$+jets & $\\ell\\ell$ & alljets & $\\ell$+jets & $\\ell\\ell$ & $\\ell$+jets & $\\ell\\ell$ & alljets & Lxy & $\\ell$+jets & $\\ell\\ell$ & MEt \\\\\n\\hline\n\n\n\nPull & $+0.40$ & $-0.51$ & $+1.11$ & $+1.32$ & $-0.38$ & $-0.51$ & $-0.82$ & $-0.41$ & $-0.67$ & $1.42$ & $+0.30$ & $+0.45$ \\\\\nWeight [\\%] & $-4.7$ & $-1.1$ & $-0.9$ & $+0.4$ & $-0.2$ & $+62.0$ & $-0.3$ & $+10.5$ & $+0.22$ & $+20.6$ & $+1.4$ & $+11.9$ \\\\\n\\hline \\hline\n\\end{tabular}\n}\n\\end{center}\n\\label{tab:stat} \n\\end{table} \n\nNo input has an anomalously large pull. \nIt is, however, still\ninteresting to determine the top-quark mass separately \nin the alljets, $\\ell$+jets, $\\ell\\ell$, and MEt channels (leaving out\nthe $\\Lxy$ measurement).\nWe use the same methodology,\ninputs, uncertainty categories, and correlations as described above, but fit \nthe four physical observables, \\ensuremath{\\MT^{\\mathrm{alljets}}}, \\ensuremath{\\MT^{\\mathrm{\\ell\\mbox{+}jets}}}, \\ensuremath{\\MT^{\\mathrm{\\ell\\ell}}}, and \\ensuremath{\\MT^{\\mathrm{MEt}}}\\ separately.\nThe results of these combinations are shown in\nFigure~\\ref{fig:three_observables} and Table~\\ref{tab:three_observables}.\n\n\\begin{figure}[p]\n\\begin{center}\n\\includegraphics[width=0.8\\textwidth]{figures\/TevMtWin13-4channels.pdf}\n\\end{center}\n\\caption[Mtop in each channel]{Summary of the combination of the 12\ntop-quark measurements by CDF and D\\O\\ for different final states.}\n\\label{fig:three_observables} \n\\end{figure}\n\n\n\nUsing the results of Table~\\ref{tab:three_observables} \n we calculate the following chi-squared values including correlations:\n$\\chi^{2}(\\ell+{\\rm jets}-\\ell\\ell)=1.30\/1$, $\\chi^{2}(\\ell+{\\rm\n jets}-{\\rm alljets})=0.07\/1$, $\\chi^{2}(\\ell+{\\rm jets}-{\\rm MEt})=0.11\/1$, \n$\\chi^{2}(\\ell\\ell-{\\rm alljets})=0.42\/1$,\n $\\chi^{2}(\\ell\\ell-{\\rm MEt})=1.22\/1$, and $\\chi^{2}({\\rm\n alljets}-{\\rm MEt})=0.19\/1$. \nThese correspond to chi-squared probabilities \nof 25\\%, 79\\%, 74\\%, 52\\%, 27\\%, and 66\\% respectively, indicating\nthat the top-quark mass determined in each decay channel is consistent\nin all cases.\n\n\\begin{table}[t]\n\\caption[Mtop in each channel]{Summary of the combination of the 12\nmeasurements by CDF and D\\O\\ in terms of four physical quantities,\nthe mass of the top quark in the alljets, $\\ell$+jets, $\\ell\\ell$, and MEt decay channels. }\n\\begin{center}\n\\renewcommand{\\arraystretch}{1.30}\n\\begin{tabular}{ccrrrr}\n\\hline\\hline\n\n\nParameter & Value (\\ensuremath{\\mathrm{ Ge\\kern -0.1em V }\\kern -0.2em \/c^2 }) & \\multicolumn{4}{c}{Correlations} \\\\\n & & $\\ensuremath{\\MT^{\\mathrm{alljets}}}$ & $\\ensuremath{\\MT^{\\mathrm{\\ell\\mbox{+}jets}}}$ & $\\ensuremath{\\MT^{\\mathrm{\\ell\\ell}}}$ & $\\ensuremath{\\MT^{\\mathrm{MEt}}}$ \\\\ \\hline\n$\\ensuremath{\\MT^{\\mathrm{alljets}}}$ & $172.7\\pm 1.9$ & 1.00 & & & \\\\\n$\\ensuremath{\\MT^{\\mathrm{\\ell\\mbox{+}jets}}}$ & $173.2\\pm 0.9$ & 0.25 & 1.00 & &\\\\\n$\\ensuremath{\\MT^{\\mathrm{\\ell\\ell}}}$ & $171.0\\pm 2.1$ & 0.19 & 0.41 & 1.00 & \\\\\n$\\ensuremath{\\MT^{\\mathrm{MEt}}}$& $173.8\\pm 1.8$ & 0.13 & 0.26 & 0.18 & 1.00 \\\\\n\\hline\\hline\n\\end{tabular}\n\\end{center}\n\\label{tab:three_observables}\n\\end{table}\n\nTo test the influence of the choices in modeling the\ncorrelations, we performed a cross-check by changing all non-diagonal\ncorrelation \ncoefficients of the correlation matrix defined in Section~\\ref{sec:corltns} from 100\\% to 50\\% and re-evaluated the combination. \nThe result of this large variation of degree of correlation is a\n$\\gevcc{+0.19}$ shift of the top-quark mass and reduces the total\nuncertainty negligibly.\nThe chosen approach is therefore conservative.\n\n\n\n\nWe also performed separate combinations of all the CDF and D\\O\\ measurements. The results of these combinations are\n$\\gevcc{172.72~\\pm~0.93}$ for CDF and $\\gevcc{174.89~\\pm~1.42}$ for D\\O. Taking all correlations into account, \nwe calculate the chi-square value $\\chi^{2}(CDF-D\\O)=2.25\/1$ corresponding to a probability of 13\\%.\n\n\\section{Summary}\n\\label{sec:summary}\n\nAn update of the combination of measurements of the mass of the top quark\nfrom the Tevatron experiments CDF and D\\O\\ has been presented. This preliminary\ncombination includes five published \\hbox{Run\\,I}\\ measurements, six published\n\\hbox{Run\\,II}\\ measurements, and \none preliminary \\hbox{Run\\,II}\\ measurement, but the majority of these\nmeasurements are not yet performed on the full datasets available. Taking into\naccount the statistical and systematic uncertainties and their\ncorrelations, the preliminary result for the Tevatron average is\n $\\ensuremath{M_{\\mathrm{t}}}=\\gevcc{\\measStatSyst{173.20}{0.51}{0.71}}$,\nwhere the total uncertainty is obtained assuming Gaussian systematic uncertainties.\nThe central value is 0.02\\,GeV\/$c^2$ higher than our July 2012\naverage~\\cite{TeVTopComboPRD} of \n$\\ensuremath{M_{\\mathrm{t}}}=173.18\\pm0.94$\\,GeV\/$c^2$.\nAdding in quadrature the statistical and systematic uncertainties\nyields a total uncertainty of $\\gevcc{0.87}$ which represents an\nimprovement of $8\\%$.\n\nThe mass of the top quark is now known with a relative precision of\n0.50\\%, limited by the systematic uncertainties, which are dominated by\nthe jet energy scale uncertainty. \nThis result will be further improved when all analysis channels from \nCDF and D\\O\\ using the full Run~II data set are finalized.\n \n\n\\section{Acknowledgments}\n\\label{sec:ack}\n\nWe thank the Fermilab staff and the technical staffs of the\nparticipating institutions for their vital contributions. \nThis work was supported by \nDOE and NSF (USA),\nCONICET and UBACyT (Argentina), \nCNPq, FAPERJ, FAPESP and FUNDUNESP (Brazil),\nCRC Program, CFI, NSERC and WestGrid Project (Canada),\nCAS and CNSF (China),\nColciencias (Colombia),\nMSMT and GACR (Czech Republic),\nAcademy of Finland (Finland),\nCEA and CNRS\/IN2P3 (France),\nBMBF and DFG (Germany),\nMinistry of Education, Culture, Sports, Science and Technology (Japan), \nWorld Class University Program, National Research Foundation (Korea),\nKRF and KOSEF (Korea),\nDAE and DST (India),\nSFI (Ireland),\nINFN (Italy),\nCONACyT (Mexico),\nNSC(Republic of China),\nFASI, Rosatom and RFBR (Russia),\nSlovak R\\&D Agency (Slovakia), \nMinisterio de Ciencia e Innovaci\\'{o}n, and Programa Consolider-Ingenio 2010 (Spain),\nThe Swedish Research Council (Sweden),\nSwiss National Science Foundation (Switzerland), \nFOM (The Netherlands),\nSTFC and the Royal Society (UK),\nand the A.P. Sloan Foundation (USA).\n\n\\clearpage\n\n\\providecommand{\\href}[2]{#2}\\begingroup\\raggedright","meta":{"redpajama_set_name":"RedPajamaArXiv"}}