diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzfdww" "b/data_all_eng_slimpj/shuffled/split2/finalzzfdww" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzfdww" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nEntanglement and its generation, besides its intrinsic interest as an unique quantum correlation and resource for quantum information processing \\cite{HHHH2009, NieChu2010}, continues to be vigorously investigated due to its relevance to a wide range of questions such as\nthermalization and the foundations of statistical physics \\cite{Popescu06,JC05,DeutchLiSharma2013,Kaufman16,Rigol2016}, decoherence \\cite{Zurek91,Haroche1998,Zurek03,Davidovich_2016}, delocalization \\cite{WSSL2008,kim13} and quantum chaos in few and many-body systems \\cite{Miller99, Bandyopadhyay02, Wang2004,LS05,petitjean2006lyapunov, tmd08, Amico08,Chaudhary}. Some of these issues\nconcern the rate of entanglement production, the nature of multipartite entanglement sharing and distribution, and long-time saturation or indefinite growth. Details of entanglement production in integrable {\\it versus} nonintegrable systems is an\nactive area of study, and it is generally appreciated that concomitant with the production of\nnear random states of nonintegrable systems is the production of large entanglement that can lead to thermalization\nof subsystems.\n\nEntanglement produced from suddenly joining two spin chains each in its ground state produces entanglement growth $\\sim \\ln(t\/a)$ at long times~\\cite{Calabrese_2007}, whereas quenches starting from arbitrary states can produce large entanglement including a linear growth phase \\cite{JC05}. Similarly in an ergodic or eigenstate thermalized phase a system can show ballistic entanglement growth \\cite{kim13,Ho17}, whereas in the many-body localized phase\nit is known to have a logarithmic growth in time. The entanglement in almost all these studies relates to bipartite block entanglement between\nmacroscopically large subsystems in many-body systems.\n\nEven earlier works have explored the entanglement in Floquet or periodically forced systems such as the coupled kicked tops or standard maps, as a means to study the relationship between chaos and entanglement \\cite{FNP1998,Miller99,Lakshminarayan2001, Bandyopadhyay02,Fujisaki03,Bandyopadhyay04}.\nIt was seen that chaos in general increases bipartite entanglement and results in near\nmaximal entanglement as the states become typical, in the sense of Haar or uniform measure.\nThe initial states considered were mostly phase-space localized coherent states. In the case when the uncoupled systems are chaotic, and the interactions are weak, after a short Ehrenfest time scale\nthe growth of the entropy or entanglement is essentially as if the coherent states were initially random subsystem states. In cases in which the subsystems are fully chaotic,\nthe growth of entanglement (beyond the Ehrenfest time) is dependent on the coupling strength rather than on measures of chaos such as the Kolmogorov-Sinai entropy or the Lyapunov exponent \\cite{Fujisaki03,Bandyopadhyay04} . A linear growth was observed in this case and perturbation theory \\cite{Fujisaki03} is successful in describing it, and the more extended time behavior was described by perturbation theory along with random matrix theory (RMT) \\cite{Bandyopadhyay04}. The linear growth leads to saturation values that are interaction independent, and in cases of moderately large coupling this is just the bipartite entanglement of typical or random states in the full Hilbert space.\n\nIn sharp contrast is the case for which the initial states are eigenstates of the non-interacting fully chaotic systems. This presents a very different scenario for weak couplings that to our knowledge has not been previously studied. The present paper develops a full theory in this case, starting\nfrom a properly regularized perturbation theory wherein a universal growth curve involving a suitably scaled time is derived. Importantly, this is further developed into a theory valid for non-perturbative strong couplings. We study these cases as an ensemble average over all uncoupled eigenstates, which clearly forms a special set of states in the Hilbert space for weak coupling.\n\n The entanglement production starts off linearly, as in the case of generic states, before saturating at much smaller entanglement values that are manifestly and strongly interaction dependent and reflect a ``memory\" of the initial ensemble to which it belongs. Interestingly while the linear regime is independent of whether the system possesses time-reversal symmetry or not, the subsequent behavior including the saturation value of the entanglement is larger for the case when time-reversal is broken. For generic initial states time-reversal symmetry has not played a significant\n role in the entanglement production or saturation.\nAs the interaction is increased this saturation approaches that of random or typical values and then the memory of starting off as a special initial state no longer persists. An essential aspect of this study is to elucidate at what interaction strength such a transition happens.\n\nThe interaction is properly measured by a scaled dimensionless transition parameter $\\Lambda$ that also determines transitions in the spectral statistics and eigenfunction entanglements of such systems \\cite{Srivastava16,Lakshminarayan16,Tomsovic18c}. If $\\Lambda=0$, the noninteracting case, although the two subsystems are chaotic, the quantum spectrum of the system has a Poisson level spacing statistic \\cite{Tkocz12} and therefore has many nearly degenerate levels which start to mix when the subsystems are weakly coupled. If $\\Lambda \\ll 1$, we are in a perturbative regime wherein the eigenstates with appreciable entanglements have a Schmidt rank of approximately two, i.e.\\ the reduced density matrix of the eigenstates has at most two principal nonzero eigenvalues. This circumstance carries over to time evolving states which are initially product eigenstates. Universal features of the eigenstate entanglement depend only on the single scaled parameter $\\Lambda$. For example, the linear entropy of the eigenstates $\\sim \\sqrt{\\Lambda}$ \\cite{Lakshminarayan16,Tomsovic18c}. On the other hand, as shown in this paper, time evolving states develop a linear entropy $\\sim C(2,t)\\sqrt{\\Lambda}$, with $C(2,t)$ being a {\\it universal} function for a properly scaled time, independent of the details of the interaction or the chaotic subsystem dynamics, except for a slight dependence on whether the system is time-reversal symmetric or not.\n\n\nSuch a universality follows from the existence of underlying RMT models that\ndescribe the transition from uncoupled to strongly coupled systems, a RMT transition ensemble. Although it is standard to\napply RMT for stationary states and spectral statistics \\cite{Bohigas84, Brody81, Haake10}, as indeed done for strongly chaotic and weakly\ninteracting systems \\cite{Srivastava16,Lakshminarayan16,Tomsovic18c}, it is noteworthy that this is typically not valid for the time evolution, because RMT lacks correlations required for describing short time dynamics properly. However the time scales\nover which the entanglement develops is much longer than the Ehrenfest time after which specific dynamical system features disappear. Thus, universal behaviors can be derived from such RMT transition ensembles provided the time scales of interest remain much longer than the Ehrenfest time scale. It turns out that for $\\Lambda \\gtrsim 1$, the interaction is strong enough that the\nsystem has fluctuations that are typical of RMT of over the whole space, for example the consecutive neighbor spacing of eigenvalues is that of Wigner \\cite{Srivastava16}. This also signals the regime for which eigenstates have typical entanglement of random states \\cite{Lakshminarayan16,Tomsovic18c} and as shown here, the time-evolving states lose memory of whether they initially belonged to special ensembles such as the noninteracting eigenstates.\n\nAlthough regularized perturbation theory, initially developed for studying symmetry breaking in\nstatistical nuclear physics~\\cite{French88a,Tomsovicthesis}, is used in the $\\Lambda \\ll 1$ regime, a novel recursive use of the perturbation theory allows for approximate, but very good extensions to the non-perturbative regime. In fact, it covers well the full transition. This provides an impressive connection of the entanglement both as a function of time and as a function of the interaction strength to the RMT regime where nearly maximal entanglement is obtained and formulas such as Lubkin's for the linear entropy \\cite{Lubkin78} and Page's for the von Neumann entropy \\cite{Page93,Sen1996} are obtained.\nWe illustrate the general theory by specifically considering both time-reversal symmetric and violating RMT transition\nensembles, given respectively by subsystem Floquet operators chosen from the circular orthogonal ensemble (COE)\nand the circular unitary ensemble (CUE) respectively. These are classic RMT ensembles consisting of unitary\nmatrices that are uniformly chosen with densities that are invariant under orthogonal (COE) and unitary (CUE) groups \\cite{MehtaBook}.\n\nIn addition, we apply it to a dynamical system of coupled standard maps \\cite{Froeschle71,Lakshminarayan2001,Richter14}. The standard map is a textbook example of a chaotic Hamiltonian system and is simply a periodically kicked pendulum. There is a natural translation symmetry in\nthe angular momentum that makes it possible to consider the classical map on a torus phase space, with periodic boundary\nconditions in both position and momentum. This yields convenient finite dimensional models of quantum chaos, the dimension\nof the Hilbert space being the inverse scaled Planck constant. The model we consider is that of coupling two\nsuch maps which has proven useful in previous studies relating to entanglement \\cite{Lakshminarayan2001,Lakshminarayan16,Tomsovic18c}, spectral transitions \\cite{Srivastava16} and out-of-time-ordered correlators \\cite{RaviLak2019}.\n\nTwo possible concrete examples are\na pair of particles in a chaotic quantum dot with tunable interactions or two spin chains that are in the ergodic phase before being suddenly joined.\nRecent experiments have accessed information of time-evolving states of interacting few-body systems via\nstate tomography of single or few particles, facilitating the study of the role of entanglement in the\napproach to thermalization of closed systems.\nSpecifically, experiments that have\nstudied nonintegrable systems include, for example, few qubit kicked top implementations \\cite{Chaudhary,Neill2016} and the Bose-Hubbard Hamiltonian \\cite{Kaufman16}. Thus modifications of these to accommodate weakly interacting parts are conceivable. This work adds to the already voluminous contemporary research on thermalization in closed systems\nby looking in detail at the time evolution for a case when thermalization, in the sense of a typical subsystem entropy, is unlikely to occur \\cite{GWEG2018}, namely starting from product eigenstates and quenching the interactions by suddenly turning them on.\n\n\nThis paper is organized as follows:\nIn Sec.~\\ref{sec:background}\nthe necessary background material on entanglement in bipartite systems,\nthe random matrix transition ensemble and the unversal\ntransition parameter are given.\nSec.~\\ref{sec:universal-entanglment-dynamics-perturbation-regime}\nprovides the perturbation theory for the\nuniversal entanglement based on the eigenvalues of the reduced\ndensity matrix, ensemble averaging for the CUE or COE\nand invoking a regularization.\nBased on this the eigenvalue moments of the reduced density matrix\nare obtained in Sec.~\\ref{sec:eigenvalue-moments}\nleading to explicit expressions for the HCT entropies.\nIn particular it is shown that for small interaction\nbetween the subsystems the simultaneous\nre-scaling of time and of the entropies by their saturation\nvalues leads to a universal curve which is independent\nof the interaction.\nThe extension to the non-perturbative regime\nis done in Sec.~\\ref{sec:NonPerturbative}\nby using a recursively embedded perturbation theory\nto produce the full transition and the saturation values.\nA comparison with a dynamical system given by\na pair of coupled kicked rotors is done\nin Sec.~\\ref{sec:coupled-kicked-rotors}.\nFinally, a summary and outlook is given in\nSec.~\\ref{sec:summary-and-outlook}.\n\n\n\n\\section{Background}\n\\label{sec:background}\n\n\\subsection{Entanglement in bipartite systems}\n\nConsider pure states, $|\\psi\\rangle$, of a bipartite system whose Hilbert space is a tensor product space, $\\mathcal{H}^A \\otimes \\mathcal{H}^B$, with subsystem dimensionalities, $N_A$ and $N_B$, respectively. Without loss of generality, let $N_A \\leq N_B$. The question to be studied is how much an initially unentangled state becomes entangled under evolution of some dynamics as a function of time.\n\nThe dynamics of a generic conservative system\ncould be governed by a Hamiltonian or by a unitary Floquet operator.\nSpecifically, a bipartite Hamiltonian system is of the form\n\\begin{equation}\nH(\\epsilon) = H_A \\otimes \\mathds{1}_B + \\mathds{1}_A \\otimes H_B + \\epsilon V_{AB} \\ , \\label{eq:GenericHamiltonian}\n\\end{equation}\nwhere the non-interacting limit is $\\epsilon = 0$. In the case of a quantum map, the dynamics can be described by a unitary Floquet operator~\\cite{Srivastava16}\n\\begin{equation}\n\\mathcal{U}(\\epsilon) = (U_A \\otimes U_B) U_{AB}(\\epsilon)\\ ,\n\\label{eq:GenericFloquet}\n\\end{equation}\nfor which the non-interacting limit is $U_{AB}(\\epsilon\\rightarrow 0) = \\mathds{1}$. We assume that both $V_{AB}$ and $U_{AB}(\\epsilon \\neq 0)$ are entangling interaction operators~\\cite{Lakshminarayan16}.\n\nThe Schmidt decomposition of a pure state is given by\n\\begin{equation}\n \\ket{\\psi} = \\sum_{l=1}^{N_A} \\sqrt{\\lambda_l} \\, \\ket{\\phi^A_l}\\ket{\\phi^B_l}.\n \\label{eq:GenericSchmidtDecomposition}\n\\end{equation}\nThe normalization condition on the state $\\ket{\\psi}$ gives\n\\begin{equation} \\label{eq:lambda-l-normalization}\n \\sum_{l=1}^{N_A} \\lambda_l = 1 .\n\\end{equation}\nThe state is unentangled if and only if the largest eigenvalue $\\lambda_1 = 1$ (all others vanishing), and maximally entangled if $\\lambda_l = 1\/N_A$ for all $l$. By partial traces, it follows that the reduced density matrices\n\\begin{equation}\n\\rho^A = \\tr_B(\\ket{\\psi}\\bra{\\psi}), \\qquad \\rho^B = \\tr_A(\\ket{\\psi}\\bra{\\psi})\\ ,\n\\end{equation}\nhave the property\n\\begin{equation}\n\\rho^A \\ket{\\phi^A_l} = \\lambda_l \\ket{\\phi^A_l}, \\quad \\text{and} \\quad \\rho^B \\ket{\\phi^B_l} = \\lambda_l \\ket{\\phi^B_l},\n\\end{equation}\n respectively. They are positive semi-definite, share the same non-vanishing (Schmidt) eigenvalues $\\lambda_l$ and $\\{ \\ket{\\phi^A_l} \\}$, and $\\{ \\ket{\\phi^B_l} \\}$ form orthonormal basis sets in the respective Hilbert spaces. For subsystem $B$ there are $N_B-N_A$ additional vanishing eigenvalues and associated eigenvectors.\n\nA very useful class of entanglement measures are given by the von Neumann entropy and Havrda-Charv\\'at-Tsallis (HCT) entropies~\\cite{Bennett96,Havrda67,Tsallis88,Bengtsson06}. The von Neumann entropy is given by\n\\begin{equation} \\label{eq:GenericVonNeumannEntropy}\n\\begin{split}\nS_1 & = -\\tr_A(\\rho^A \\ln \\rho^A ) = -\\tr_B(\\rho^B \\ln \\rho^B )\\\\\n &= -\\sum_{l=1}^{N_A} \\lambda_l \\ln \\lambda_l,\n\\end{split}\n\\end{equation}\nwhich vanishes if the state is unentangled and is maximized if all the nonvanishing eigenvalues are equal to $1\/N_A$. The HCT entropies are obtained from moments of the Schmidt eigenvalues. Defining\n\\begin{equation}\n\\mu_\\alpha = \\tr_A[(\\rho^A)^\\alpha] = \\tr_B[(\\rho^B)^\\alpha] = \\sum_{l=1}^{N_A} \\lambda_l^\\alpha, \\quad \\alpha >0 \\ ,\n\\label{eq:GenericMoments}\n\\end{equation}\ngives the HCT entropies as\n\\begin{equation} \\label{eq:GenericHCTEntropy}\n S_\\alpha = \\frac{1-\\mu_\\alpha}{\\alpha-1}.\n\\end{equation}\nNote that these differ from the R\\'enyi entropies \\cite{Ren1961wcrossref},\nwhich are defined by\n\\begin{equation*}\n R_\\alpha = \\dfrac{\\ln \\mu_{\\alpha}}{1-\\alpha} .\n\\end{equation*}\nIn the limit $\\alpha \\to 1$ also $R_\\alpha$ turns into the\nvon Neumann entropy.\nIn this work we use the HCT entropies as performing ensemble averages\nis easier using $\\mu_{\\alpha}$ than $\\ln \\mu_{\\alpha}$.\n\n\n\n\\subsection{Quantum chaos, random matrix theory, and universality}\n\nMany statistical properties of strongly chaotic quantum systems are successfully modeled and derived with the use of RMT~\\cite{Brody81,Bohigas84}. Generally speaking, the resulting properties are universal, and in particular, do not depend on any of the physical details of the system with the exception of symmetries that it respects. Here the subsystems are assumed individually to be strongly chaotic. Thus, the statistical properties of the dynamics, Eq.~(\\ref{eq:GenericFloquet}), can be modeled with the operators $U_A$ and $U_B$ being one of the standard circular RMT ensembles~\\cite{Dyson62e}, orthogonal, unitary, or symplectic, depending on the fundamental symmetries of the system~\\cite{Porterbook}.\n\nWe concentrate on the orthogonal (COE) and unitary ensembles (CUE) depending on whether or not time reversal invariance is preserved, respectively.\nThe derivation of the typical entanglement production for some initial state relies on the dynamics governed by\nthe random matrix transition ensemble \\cite{Srivastava16,Lakshminarayan16}\n\\begin{equation}\n\\mathcal{U}_{\\text{RMT}}(\\epsilon) = (U_A^\\text{RMT} \\otimes U_B^\\text{RMT}) U_{AB}(\\epsilon)\\ .\n\\label{eq:GenericFloquetRMT}\n\\end{equation}\nThe operator $U_{AB}(\\epsilon)$ is assumed to be diagonal in the direct product basis of the two subsystem ensembles. Explicitly, the diagonal elements are considered to be of the form $\\exp(2 \\pi i \\epsilon \\xi_{kl})$, where $\\xi_{kl}$ ($1\\le k,l\\le N_A,N_B$) is a random number uniformly distributed in $(-1\/2,1\/2]$.\n\n\n\\subsection{Symmetry breaking and the transition parameter}\n\nThe statistical properties of weakly interacting quantum chaotic bipartite systems have been studied recently, with the focus on spectral statistics, eigenstate entanglement, and measures of localization~\\cite{Srivastava16,Lakshminarayan16,Tomsovic18c}. If the subsystems are not interacting, the spectrum of the full system is just the convolution of the two subsystem spectra giving an uncorrelated spectrum in the large dimensionality limit. The eigenstates of the system are unentangled. It is very fruitful to conceptualize this as a dynamical symmetry. Upon introducing a weak interaction between the subsystems, this symmetry is weakly broken. As the interaction strength increases, the spectrum becomes increasingly correlated, and the eigenstates entangled.\n\nHere $U_{AB}(\\epsilon)$ plays the role of a dynamical symmetry breaking operator. For $\\epsilon=0$, the symmetry is preserved (the dynamics of the subsystems are completely independent), and as $\\epsilon$ gets larger, the more complete the symmetry is broken. It is known that for sufficiently chaotic systems, there is a universal scaling given by a transition parameter which governs the influence of the symmetry breaking on the system's statistical properties~\\cite{French88a}. The transition parameter is defined as~\\cite{Pandey83}\n\\begin{equation}\n\\Lambda = \\frac{v^2(\\epsilon)}{D^2}, \\label{eq:LambdaDef}\n\\end{equation}\nwhere $D$ is the mean level spacing and $v^2(\\epsilon)$ is the mean square matrix element in the eigenbasis of the symmetry preserving system, calculated locally in the spectrum.\n\nFor the COE and CUE the leading behavior in $N_A$ and $N_B$ is \\cite{Srivastava16,Tomsovic18c,HerKieFriBae2019:p}\n\\begin{align}\n\\label{eq:Lambda-RMT}\n& \\, \\Lambda = \\frac{N_A N_B}{4 \\pi^2} \\left[1-\\dfrac{\\sin^2 (\\pi \\epsilon)}{\\pi^2\\epsilon^2} \\right]\n\\sim \\frac{\\epsilon^2 N_AN_B}{12}\\ ,\n\\end{align}\nwhere the last result is in the limit of large $N_A$, $N_B$.\nThe transition parameter $\\Lambda$ ranges over $0 \\le \\Lambda \\le N_A N_B\/4\\pi^2\\ (N_A, N_B \\rightarrow \\infty)$, where the limiting cases are fully symmetry preserving, and fully broken, respectively.\nIn essence, the latter expression of Eq.~\\eqref{eq:Lambda-RMT} illustrates the fact that as the system size grows, a symmetry breaking transition has a discontinuously fast limit in $\\epsilon$.\n\nThe transition parameter gives the relation necessary to compare the statistical properties of systems of any size and kind to each other. As long as $\\Lambda$ has identical values, the systems have identical properties. However, for a particular dynamical system, it can turn out to be rather difficult to calculate $\\Lambda$. Although, the statistical properties are universal and independent of the nature of the system in this chaotic limit, properties such as whether the system is many-body or single particle, Fermionic or Bosonic, actually enter into its calculation. For example, a method for calculating $\\Lambda$ for highly excited heavy nuclei is given in Ref.~\\cite{French88b}. The far simpler case of coupled kicked rotors is given in Ref.~\\cite{Srivastava16}, and is used ahead for illustration. In extended systems, the issue of localization emerges, which must also be taken into account, and for them the term sufficiently chaotic is meant to exclude a localized regime.\n\n\\section{Universal entanglement production -- Perturbative regime}\n\\label{sec:universal-entanglment-dynamics-perturbation-regime}\n\nThe starting point of a derivation of the typical production rate of entanglement in initially unentangled states is the random matrix transition ensemble~\\eqref{eq:GenericFloquetRMT}. Following a similar derivation sequence for the eigenstates in Refs.~\\cite{Lakshminarayan16,Tomsovic18c}, the first step is to derive expressions for the eigenvalues of the reduced density matrix, which can be obtained from the Schmidt decomposition of the time evolved state of the system. Applying a standard Rayleigh-Schr{\\\"o}dinger perturbation theory leads to perturbation expressions for the Schmidt eigenvalues. However, due to the Poissonian fluctuations in the spectrum of the non-interacting system, near-degeneracies occur too frequently and cause divergences in the ensemble averages. It is therefore necessary first to regularize the eigenvalue expressions appropriately. It also turns out that the perturbation expressions for the HCT entanglement measures can be further extended to a non-perturbative regime by recursively invoking the regularized perturbation theory leading to a differential equation, which is analytically solvable~\\cite{Tomsovic18c},\nsee Sec.~\\ref{sec:NonPerturbative}.\n\n\n\\subsection{Definitions}\n\nThe eigenvalues and corresponding eigenstates of the unitary operators\n$U_A$ and $U_B$ for the subsystems\nand of $\\mathcal{U}(\\epsilon) = (U_A \\otimes U_B) U_{AB}(\\epsilon)$\nof the full bipartite system \\eqref{eq:GenericFloquet} are given\nby the equations\n\\begin{eqnarray}\nU_A \\ket{j^A} &=& \\text{e}^{i \\theta_j^A} \\ket{j^A},\\quad j=1,2,3,\\ldots,N_A \\nonumber \\\\\nU_B \\ket{k^B} &=& \\text{e}^{i \\theta_k^B} \\ket{k^B}, \\quad k=1,2,3,\\ldots,N_B\\\\\n\\mathcal{U}(\\epsilon) \\ket{\\Phi_{jk}} &=& \\text{e}^{i \\varphi_{jk}} \\ket{\\Phi_{jk}}\\ .\n\\nonumber\n\\end{eqnarray}\nTo simplify the notation, the superscripts $A$ and $B$ are dropped for both eigenkets, $\\ket{j^A}\\ket{k^B} \\equiv \\ket {jk}$, and the eigenvalues $\\theta_j^A \\equiv \\theta_j$ ($\\theta_k^B \\equiv \\theta_k$). It is understood that the labels $j$ and $k$ are reserved for the subsystems $A$ and $B$, respectively. Similarly for convenience, the subscript $AB$ is dropped from the operator $V_{AB}$.\n\nGiven the form \\eqref{eq:GenericFloquet} of the unitary operator $\\mathcal{U}(\\epsilon)$, in the limit $\\epsilon \\rightarrow 0$ one has $\\ket{\\Phi_{jk}} \\rightarrow \\ket{jk}$ which is a product eigenstate of the unperturbed system and forms a complete basis with spectrum $\\varphi_{jk} \\rightarrow \\theta_{jk} = \\theta_j + \\theta_k \\,\\,\\text{mod}\\,\\, 2\\pi$. For non-vanishing $\\epsilon$ there is a unitary transformation $S$ between the eigenbases for the set $\\ket{\\Phi_{jk}}$ and $\\ket{jk}$ whose matrix elements can be identified using the relations\n\\begin{eqnarray}\n\\ket{\\Phi_{jk}} &=& \\sum_{j'k'} S_{jk,j'k'} \\ket{j'k'} = \\sum_{j'k'} \\ket{j'k'} \\bra{j'k'}\\ket{\\Phi_{jk}} \\nonumber \\\\\n\\ket{jk} &=& \\sum_{j'k'} S^\\dagger_{jk,j'k'} \\ket{\\Phi_{j'k'}}\\ .\n\\label{eq:UnitaryTransform}\n\\end{eqnarray}\n\n\\subsection{Eigenvalues of the reduced density matrix}\n\nIn the limit $N_A \\rightarrow \\infty$, perturbation theory for unitary Floquet systems generates the same equations as for Hamiltonian systems up to vanishing corrections of $\\mathcal{O}((N_A N_B)^{-1})$ if one identifies $U_{AB}(\\epsilon)=\\exp(i\\epsilon V)$~\\cite{Tomsovic18c}. For an initial unentangled state, begin by considering an eigenstate $\\ket{jk}$ of the non-interacting system. Denote the time evolution of this initial state after $n$ iterations of the dynamics as $\\ket{j k(n;\\epsilon)}$ [$ = \\mathcal{U}^n(\\epsilon) \\ket{j k}$]. Upon the usual insertion of the completeness relation one gets\n\\begin{equation}\n\\ket{jk(n;\\epsilon)} = \\sum_{j'k'} \\text{e}^{i n \\varphi_{j'k'}} S^\\dagger_{jk,j'k'} \\ket{\\Phi_{j'k'}}. \\label{eq:GeneralTimeEvolvedState}\n\\end{equation}\nThis time evolved state has a standard Schmidt decomposed form\n\\begin{equation}\n\\ket{jk(n;\\epsilon)} = \\sum_{l = 1}^{N_A} \\sqrt{\\lambda_l(n;\\epsilon)} \\, \\ket{\\phi^A_l(n;\\epsilon)}\\ket{\\phi^B_l(n;\\epsilon)},\n\\label{eq:GeneralTimeEvolvedStateInSDForm}\n\\end{equation}\nwhere $\\lambda_1 \\geq \\lambda_2 \\geq \\ldots \\geq \\lambda_{N_A}$ are time-dependent Schmidt numbers (eigenvalues of the reduced density matrices) such that $\\sum_l \\lambda_l(n;\\epsilon) = 1$, and $\\{ \\ket{\\phi^A_l(n;\\epsilon)}\\}$, $\\{ \\ket{\\phi^B_l(n;\\epsilon)}\\}$ are the corresponding Schmidt eigenvectors of the $A$ and $B$ subspaces, respectively.\n\nIt was shown in Ref.~\\cite{Lakshminarayan16} that for weak perturbations, the Schmidt decomposition of the eigenstates to $\\mathcal{O}(N_A^{-1})$ corrections are given by the neighboring eigenstates of the unperturbed (non-interacting) system and the perturbation theory coefficients. This can be considered as a kind of automatic Schmidt decomposition.\n The generalization to the time evolving states, $\\ket{j k(n;\\epsilon)}$, follows by another insertion of the unitary transformation $S$ to give\n\\begin{equation}\n\\ket{jk(n;\\epsilon)} = \\sum_{j''k''} \\sum_{j'k'} \\text{e}^{i n \\varphi_{j'k'}} S^\\dagger_{jk,j'k'} S_{j'k',j''k''} \\ket{j''k''}. \\label{eq:GeneralTimeEvolvedState2}\n\\end{equation}\nThis leads to the identification\n\\begin{equation}\n\\lambda_l(n;\\epsilon) = \\left| \\sum_{j'k'} \\text{e}^{i n \\varphi_{j'k'}} S^\\dagger_{jk,j'k'} S_{j'k',(jk)_l} \\right|^2,\\label{eq:SchmidtNumbersPT}\n\\end{equation}\nwhere $j''k''\\rightarrow (jk)_l$, meaning that fixing $l$ fixes a unique and distinct index pair $(jk)_l$; e.g.~$(jk)_1=jk$. Only a small subset ($\\lesssim N_A$) of possible pairs $j''k''$ are related to a $(jk)_l$ due to the energy denominators in perturbation theory. This is a direct result of the automatic Schmidt decomposition.\nFor $\\epsilon = 0$ one has $\\lambda_1(n;0) =1$, and the rest of the Schmidt eigenvalues vanish\nby the normalization \\eqref{eq:lambda-l-normalization}, as the initial state\nis a product state of eigenstates of the two subsystems.\n\nTo prepare for ensemble averaging, it is helpful to: i) assume that the $jk$ pairs are ordered by the order of the eigenvalues $\\varphi_{jk}$, ii) use the properties of $S$ so that the $n=0$ results are immediately evident, and iii) separate out the diagonal matrix element $S_{jk,jk}$ as a special case. Let $\\Delta \\varphi_{jk,j^\\prime k^\\prime} = \\varphi_{jk} - \\varphi_{j^\\prime k^\\prime}$. For the largest eigenvalue, i.e.~Eq.~(\\ref{eq:SchmidtNumbersPT}) for $l = 1$, one finds\n\\begin{align}\n& \\lambda_1(n;\\epsilon) = 1 - 2 \\sum_{j'k', j''k''} \\left|S_{j'k',jk}\\right|^2 \\left|S_{j''k'',jk}\\right|^2 \\nonumber \\\\\n& \\qquad \\qquad \\qquad \\qquad \\qquad \\quad \\times \\sin^2 \\left(\\frac{n \\Delta \\varphi_{j'k',j''k''}}{2} \\right) \\nonumber \\\\\n& = 1 - 4 \\left|S_{jk,jk}\\right|^2\\sum_{j'k'} \\left|S_{j'k',jk}\\right|^2 \\sin^2 \\left(\\frac{n \\Delta \\varphi_{jk,j'k'}}{2} \\right) \\nonumber \\\\\n& - 4 \\sum_{\\substack{j'k' \\le j''k'' \\\\ \\ne jk}} \\left|S_{j'k',jk}\\right|^2 \\left|S_{j''k'',jk}\\right|^2 \\sin^2 \\left(\\frac{n \\Delta \\varphi_{j'k',j''k''}}{2} \\right). \\nonumber \\\\\n\\label{eq:LargestEigenvalueRaw}\n\\end{align}\nFor $l \\ge 2$, and thus $(jk)_l \\neq jk$, a similar manipulation gives\n\\begin{align}\n&\\lambda_l(n;\\epsilon) = - \\sum_{\\substack{j'k'\\\\j''k''}} \\Re\\left\\{S^\\dagger_{jk,j'k'} S_{j'k',(jk)_l} S^\\dagger_{(jk)_l,j''k''} S_{j''k'',jk} \\right\\} \\nonumber \\\\\n& \\qquad \\qquad \\qquad \\qquad \\times 2 \\sin^2 \\left(\\frac{n \\Delta \\varphi_{j'k',j''k''}}{2} \\right) \\nonumber \\\\\n& \\qquad \\qquad - \\sum_{\\substack{j'k'\\\\j''k''}} \\Im\\Big\\{S^\\dagger_{jk,j'k'} S_{j'k',(jk)_l} S^\\dagger_{(jk)_l,j''k''} S_{j''k'',jk} \\Big\\} \\nonumber \\\\\n& \\qquad \\qquad \\qquad \\qquad \\times \\sin \\left(n \\Delta \\varphi_{j'k',j''k''} \\right).\n\\label{eq:OtherEigenvaluesRaw}\n\\end{align}\nNote that summing Eq.~\\eqref{eq:OtherEigenvaluesRaw}\nover $l>1$ reproduces unity minus the expression of Eq.~\\eqref{eq:LargestEigenvalueRaw} as it must. These Schmidt eigenvalue expressions are exact to order $\\mathcal{O}(N_A^{-1})$.\n\nLowest order Rayleigh-Schr\\\"{o}dinger perturbation theory is applied to the matrix elements of $S$ in order to obtain the complete $O(\\epsilon^2)$ terms of the corresponding expressions for the Schmidt eigenvalues. Let $\\Delta \\theta_{jk,j'k'} = \\theta_{jk}-\\theta_{j'k'}$. The matrix elements $S_{jk,j'k'}$ are approximately\n\\begin{equation}\nS_{jk,j'k'} \\approx\n\\begin{cases}\n\\frac{1}{\\sqrt{\\mathcal{N}_{jk}}}, & jk = j'k' \\\\\n\\frac{1}{\\sqrt{\\mathcal{N}_{jk}}} \\frac{\\epsilon \\, V_{j'k',jk}}{\\Delta \\theta_{jk,j'k'}}, & jk \\neq j'k',\n\\end{cases} \\label{eq:UnitaryTransformPT}\n\\end{equation}\nwhere $\\mathcal{N}_{jk}$ is the normalization factor and the perturbed quasienergy is\n\\begin{align}\n& \\varphi_{jk} = \\theta_{jk} + \\epsilon^2 \\sum_{j'k' \\neq jk } \\frac{|V_{j'k',jk}|^2}{\\Delta \\theta_{jk,j'k'}}. \\label{eq:Unreg_quasienergy}\n\\end{align}\nNote first that in this derivation the diagonal matrix elements $V_{j'k',j'k'}$ are set to zero, because the energy shift due to the first order correction is a random number added to an uncorrelated spectrum giving another uncorrelated spectrum, and hence will not change the spectral statistics nor rotate the eigenvectors. Secondly, the normalization factor is included even though its first correction is $O(\\epsilon^2)$ because it plays a significant role in determining the regularized expressions ahead, likewise for the perturbed eigenvalues multiplied by the time in the argument of the sine function. For the largest eigenvalue Eq.~(\\ref{eq:LargestEigenvalueRaw}) becomes\n\\begin{align}\n&\\lambda_1(n;\\epsilon) \\approx 1 - \\frac{4}{\\mathcal{N}_{jk}} \\sum_{j'k' \\neq jk} \\bigg(\\frac{\\epsilon^2 \\, |V_{jk,j'k'}|^2}{\\mathcal{N}_{j'k'} \\, \\Delta \\theta_{j'k',jk}^2} \\bigg) \\nonumber \\\\\n& \\qquad \\qquad \\qquad \\qquad \\times \\sin^2 \\left(\\frac{n \\Delta \\varphi_{j'k',jk}}{2}\\right),\n\\label{eq:LargestEigenvaluePT}\n\\end{align}\nand for $l \\ge 2$ the others, Eq.~\\eqref{eq:OtherEigenvaluesRaw}, read\n\\begin{align}\n& \\lambda_l(n;\\epsilon) \\approx \\frac{4}{\\mathcal{N}_{jk}}\n \\frac{\\epsilon^2\\,|V_{(jk)_l,jk}|^2}{\\mathcal{N}_{(jk)_l}\\,\\Delta \\theta_{jk,(jk)_l}^2} \\sin^2 \\left(\\frac{n \\Delta \\varphi_{(jk)_l,jk}}{2}\\right),\n\\label{eq:OtherEigenvaluesPT}\n\\end{align}\nwhere $\\mathcal{N}_{jk}=\\mathcal{N}_{(jk)_l}=1 + \\mathcal{O}(\\epsilon^2)$.\n\n\\subsection{Ensemble averaging}\nBefore moving on to ensemble averaging, it is helpful to make some rescalings as follows:\n\\begin{align}\n& \\Delta \\theta_{j'k',jk} = D \\, s_{j'k'}\\\\\n& \\Delta \\varphi_{j'k',jk} = D \\, s_{j'k'}(\\epsilon) \\approx D s_{j'k'} \\bigg( 1 + \\frac{2 \\Lambda w_{j'k'}}{s^2_{j'k'}} \\bigg) \\label{eq:UnregSpacing}\\\\\n& \\epsilon^2 |V_{jk,j'k'}|^2 = \\Lambda D^2 \\, w_{j'k'}\\ ,\n\\end{align}\nwhere $D = 2\\pi\/(N_A N_B)$ is the mean level spacing of the full system, $s_{j'k'}=s_{j'k'}(0)$, and the $jk$ subscript is dropped where unnecessary. The approximation in Eq.~\\eqref{eq:UnregSpacing} follows by considering only the matrix element that directly connects the two levels. The other terms in Eq.~\\eqref{eq:UnregSpacing} move the levels back and forth and mostly cancel, but this term pushes the two levels away from each other and is dominant when the two levels are close lying where the correction may contribute.\nThe symmetry breaking (entangling) interaction matrix elements of $V$ represented in the eigenbasis of the unperturbed system behave as complex Gaussian random variable such that\n$ \\overline{ | V_{jk,j'k'} |^2 } = v^2 w_{j'k'} $, where $w_{j'k'}$ follow a Porter-Thomas distribution \\cite{PorTho1956} for the COE and an exponential one for the CUE:\n\\begin{equation}\n\\rho(w) =\n\\begin{cases}\n\\frac{1}{\\sqrt{2\\pi w}}\\exp(-w\/2) & \\qquad \\text{for COE} \\\\\n\\exp(-w) & \\qquad \\text{for CUE}.\n\\end{cases}\\label{PorterThomas}\n\\end{equation}\nIn both of the cases, $\\overline{w_{j'k'} }= 1$, which is consistent with $\\Lambda = \\epsilon^2 v^2 \/D^2$. In real dynamical systems, deviations from Porter-Thomas distributions may occur as noted in Ref.~\\cite{Tomsovic18c}.\n\nThus, in the rescaled variables the Schmidt eigenvalues for $l \\ge 2$ are\n\\begin{equation}\n\\lambda_l(n;\\Lambda) \\approx \\frac{4}{\\mathcal{N}_{jk}}\\Bigg(\\frac{\\Lambda \\,w_{(jk)_l} }{\\mathcal{N}_{(jk)_l}\\, s_{(jk)_l}^2} \\Bigg) \\,\\sin^2\\bigg(\\frac{n Ds_{(jk)_l}{(\\epsilon)}}{2}\\bigg), \\label{eq:OtherEigenvaluesRescaled}\n\\end{equation}\nand the relation, following from the normalization\ncondition Eq.~\\eqref{eq:lambda-l-normalization},\n\\begin{equation}\n\\lambda_1(n;\\Lambda) = 1 - \\sum_{l\\ne 1}\\lambda_l(n;\\Lambda)\n\\end{equation}\nis exactly preserved to this order. Next convert the expressions for the Schmidt eigenvalues into integrals, by making use of the function $R(s,w)$~\\cite{Tomsovic18c},\n\\begin{equation}\nR(s,w) = \\sum_{j'k' \\neq jk} \\delta(w-w_{j'k'}) \\delta(s-s_{j'k'}), \\label{eq:R-function}\n\\end{equation}\nwhich after ensemble averaging becomes the joint probability density of finding a level at a rescaled distance $s$ from $\\theta_{jk}$ and the corresponding scaled matrix element $w_{j'k'}$ at the value $w$.\nWith these definitions, scalings, and substitutions, Eq.~(\\ref{eq:LargestEigenvaluePT}) becomes\n\\begin{align}\n& \\lambda_1(n;\\epsilon) \\approx 1 - 4 \\Lambda \\int_{-\\infty}^\\infty \\dd{s} \\int_0^\\infty \\dd{w} \\frac{w}{s^2}\\,R(s,w) \\nonumber \\\\\n&\\qquad \\qquad \\qquad \\times \\sin^2\\Big(\\frac{n D s}{2}\\big[\\,1+\\frac{2 \\Lambda w}{s^2}\\,\\big]\\Big).\n\\end{align}\nThe ensemble average of $\\lambda_1(n;\\epsilon)$ follows by substituting the ensemble average of $R(s,w)$ by\n\\begin{equation}\n\\overline{R(s,w)} = R_2(s) \\, \\rho(w),\n\\end{equation}\nwhere $R_2(s)$ is the two-point correlation function and $\\rho(w)$ is defined in Eq.~(\\ref{PorterThomas}). For an uncorrelated spectrum $R_2(s)=1$ for $-\\infty < s < \\infty$.\nTherefore, the averaged largest Schmidt eigenvalue is\n\\begin{align}\n& \\overline{ \\lambda_1(n;\\Lambda)}\\approx 1 - 4 \\Lambda \\int_{-\\infty}^\\infty \\dd{s} \\int_0^\\infty \\dd{w} \\frac{w}{s^2}\\, \\rho(w) \\nonumber \\\\\n&\\qquad \\qquad \\qquad \\times \\sin^2\\Big(\\frac{n D s}{2}\\big[\\,1+\\frac{2 \\Lambda w}{s^2}\\,\\big]\\Big). \\label{eq:LargestEigenvalueUnreg}\n\\end{align}\nThis expression diverges due to the fact that too many spacings are vanishingly small across the ensemble, and the perturbation theory must account for spacings smaller than the matrix elements. In the next subsection the expressions are regularized properly for small $s$.\n\nIt is worth noting that if the interest is in the ensemble average of some function of $\\lambda_1(n;\\Lambda)$, then one must consider the ensemble average of the same function of $R(s,w)$. Perhaps, the simplest example is the ensemble average of the square of $\\lambda_1(n;\\Lambda)$ for which the needed result is~\\cite{Tomsovic18c}\n\\begin{eqnarray}\n\\overline{R(s_1,w_1)R(s_2,w_2)} &=& R_3(s_1,s_2) \\rho(w_1)\\rho(w_2) + \\nonumber \\\\\n&&\\hspace*{-2cm}\n \\delta\\left(w_1-w_2\\right)\\rho(w_1) \\delta\\left(s_1-s_2\\right)R_2(s_1)\n\\label{threepoint}\n\\end{eqnarray}\nwhich involves both the $2$-point and $3$-point spectral correlation functions. However, it turns out that the leading correction depends on $R_2(s)$, as the $R_3(s_1,s_2)$ term gives a contribution that is $\\mathcal{O}(\\sqrt{\\Lambda})$ smaller in comparison, and for example, generating the leading correction of high order moments depends only on the $2$-point spectral correlation function. This circumstance is helpful ahead in the next section.\n\nFollowing the same sequence of steps for the second largest eigenvalue $\\lambda_2$ requires, in addition, the probability density of the closest scaled energy of one of the $\\ket{(jk)_l}$. For uncorrelated spectra it is given by $\\rho_{\\text{CN}}(s) = 2 \\exp(-2 s)$ for $0 \\le s < \\infty$~\\cite{Tomsovic18c,Srivastava19}. One finds for the ensemble average of second largest eigenvalue,\n\\begin{align}\n& \\overline{\\lambda_2(n;\\Lambda)} \\approx 4 \\Lambda \\int_{-\\infty}^\\infty \\dd{s} \\int_0^\\infty \\dd{w} \\frac{w}{s^2} \\,\\rho(w) \\, \\rho_{\\text{CN}}(s)\\nonumber \\\\\n&\\qquad \\qquad \\qquad \\times \\sin^2\\Big(\\frac{n D s}{2}\\big[\\,1+\\frac{2 \\Lambda w}{s^2}\\,\\big]\\Big),\\label{eq:OtherEigenvalueUnreg}\n\\end{align}\nwhich is also divergent for small $s$. It turns out that the apparent order of corrections, $\\mathcal{O}(\\Lambda)$, seen in Eqs.~(\\ref{eq:LargestEigenvalueUnreg}, \\ref{eq:OtherEigenvalueUnreg}), is not correct. The regularization required to deal with the small energy denominators in the perturbation expressions alters the leading order to $\\mathcal{O}(\\sqrt{\\Lambda})$.\n\n\\subsection{Regularized perturbation theory}\n\nThe method for regularizing the perturbation expressions was introduced in Refs.~\\cite{Tomsovicthesis,French88a} and developed for the Schmidt eigenvalues pertaining to the eigenstates of interacting quantum chaotic systems in Refs.~\\cite{Lakshminarayan16,Tomsovic18c}. The standard Rayleigh-Schr\\\"{o}dinger perturbation expressions break down when the unperturbed spectrum has nearly degenerate levels due to the small energy denominators. However, there is an infinite sub-series of terms within the perturbation series of a quantity of interest which involve only two levels that are diverging due to near-degeneracy. This subseries can be resummed to get the corresponding regularized expressions. These results are equivalent to the two-dimensional degenerate perturbation theory results.\n\nThe regularized expressions for the Schmidt eigenvalues upon resummation of the two-level like terms of the perturbation series boil down to essentially replacing\n\\begin{align}\n\\frac{1}{\\mathcal{N}_{jk}}\\Bigg(\\frac{\\Lambda\\,w_{j'k'} }{\\mathcal{N}_{j'k'}\\, s_{j'k'}^2} \\Bigg) \\mapsto \\frac{\\Lambda \\, w_{j'k'}}{s_{j'k'}^2+ 4 \\, \\Lambda \\, w_{j'k'}} , \\label{eq:RegularizedNormTimesMatrixElement}\n\\end{align}\nalong with the energy spacing~\\cite{Srivastava16} in Eq.~(\\ref{eq:UnregSpacing}) as\n\\begin{equation}\ns_{j'k'}(\\epsilon) \\mapsto \\sqrt{s_{j'k'}^2 + 4 \\Lambda w_{j'k'} } \\label{eq:RegularizedSpacing}\n\\end{equation}\nin Eqs.~(\\ref{eq:LargestEigenvaluePT}, \\ref{eq:OtherEigenvaluesPT}). To verify the result in Eq.~(\\ref{eq:RegularizedNormTimesMatrixElement}), the Schmidt eigenvalues in Eqs.~(\\ref{eq:LargestEigenvalueRaw}, \\ref{eq:OtherEigenvaluesRaw}) were expanded using perturbation theory of the matrix elements $S_{jk,j'k'}$ up to and including order $\\mathcal{O}(\\epsilon^4)$. The details for this are given in App.~\\ref{app:RegularizationDerivation}. For a two-level system the normalization is\n\\begin{equation}\n|S_{jk,jk}|^2 = \\frac{1}{\\mathcal{N}_{jk}} = \\frac{1}{2}\\Bigg( 1 + \\frac{|s_{j'k'}|}{\\sqrt{s_{j'k'}^2+4 \\Lambda w_{j'k'}}} \\Bigg)\n\\end{equation}\nand the matrix element\n\\begin{equation}\n|S_{j'k',jk}|^2 = \\frac{1}{2}\\Bigg( 1 - \\frac{|s_{j'k'}|}{\\sqrt{s_{j'k'}^2+4 \\Lambda w_{j'k'}}} \\Bigg).\n\\end{equation}\nUsing these results, we get Eq.~(\\ref{eq:RegularizedNormTimesMatrixElement}). This gives the regularized Schmidt eigenvalues for $l \\neq 1$ as\n\\begin{align}\n& \\lambda_l(n;\\Lambda) = \\frac{4 \\, \\Lambda \\, w_{(jk)_l}}{s^2_{(jk)_l} + 4 \\, \\Lambda \\, w_{(jk)_l} } \\nonumber \\\\\n& \\qquad \\qquad \\qquad \\times \\sin^2 \\Big( \\frac{n D}{2} \\sqrt{s^2_{(jk)_l} + 4 \\, \\Lambda \\, w_{(jk)_l}} \\,\\Big). \\label{eq:SchmidtEigenvaluelneq1Reg}\n\\end{align}\nRescaling the spacing $z = s\/\\sqrt{\\Lambda}$ and time\n\\begin{equation} \\label{eq:time-rescaling}\n t = n D\\sqrt{\\Lambda} ,\n\\end{equation}\nthe ensemble average of the first two Schmidt eigenvalues is given by\n\\begin{align}\n& \\overline{ \\lambda_1(t;\\Lambda)} = 1- \\sqrt{\\Lambda}\\int_0^\\infty \\dd{w} \\int_{-\\infty}^\\infty \\dd{z} \\frac{4 w}{z^2 + 4 w} \\, \\rho(w) \\nonumber \\\\\n& \\qquad \\qquad \\qquad \\times \\sin^2\\Big(\\frac{t}{2} \\sqrt{z^2 + 4 w} \\Big) \\nonumber \\\\\n& \\qquad \\quad \\, = 1 - C_2(1;t)\\,\\sqrt{\\Lambda},\n\\end{align}\nwhere $C_2(1;t)$ is a short-hand for the integral\n(the general notation for arbitrary moments is given in\nEq.~(\\ref{eq:C2Integral}) ahead) and,\n\\begin{align}\n& \\overline{ \\lambda_2(t;\\Lambda)} = \\sqrt{\\Lambda}\\int_0^\\infty \\dd{w} \\int_0^\\infty \\dd{z} \\frac{4 w}{z^2 + 4 w} \\nonumber \\\\\n& \\qquad \\qquad \\qquad \\times \\, \\rho(w) \\,\\big(2 \\,\\text{e}^{-2 z \\sqrt{\\Lambda}}\\,\\big) \\sin^2\\Big(\\frac{t}{2} \\sqrt{z^2 + 4 w} \\Big) \\nonumber \\\\\n& \\, \\quad \\qquad = C_{2}(1;t) \\sqrt{\\Lambda} + \\mathcal{O}(\\Lambda \\ln \\Lambda).\n\\end{align}\nFor sufficiently small $\\Lambda$ it turns out that for both the COE and CUE cases, only (unperturbed) eigenstates corresponding to the first two largest Schmidt eigenvalues contribute largely to the state $\\ket{jk(t;\\Lambda)}$, i.e.,\n\\begin{equation}\n\\overline{ \\lambda_1(t;\\Lambda) + \\lambda_2(t;\\Lambda)} = 1 + \\mathcal{O}(\\Lambda \\ln \\Lambda ),\n\\end{equation}\nand other Schmidt eigenvalues ($l>2$) contribute in higher orders than $\\sqrt{\\Lambda}$.\nThis is crucial for extending the perturbation theory of the Schmidt eigenvalue moments to the non-perturbative regime, which is done in Sec.~\\ref{sec:NonPerturbative}.\nIt should be noted that as the unperturbed spectrum is uncorrelated, there is a non-zero probability of three-level, four-level and so-forth near-degeneracy occurrences, but with lower probability from the two-level case, and hence their contributions are higher of order than $\\sqrt{\\Lambda}$.\n\nMoreover note that the perturbation expressions for the Schmidt eigenvalues of the eigenstates $\\{\\ket{\\Phi_{jk}}\\}$ given in Refs.~\\cite{Lakshminarayan16, Tomsovic18c} for the largest eigenvalue and the other eigenvalues are $1 - \\sum_{j'k \\neq jk} \\epsilon^2 |V_{(j'k',jk}|^2\/\\Delta\\theta^2_{jk,j'k'}$ and $\\epsilon^2 |V_{(jk)_l,jk}|^2\/\\Delta\\theta^2_{jk,(jk)_l}$, respectively, in contrast to the expressions for the Schmidt eigenvalues of a time evolving state $\\ket{jk(n;\\epsilon)}$ presented in Eqs.~(\\ref{eq:LargestEigenvaluePT}, \\ref{eq:OtherEigenvaluesPT}). Due to an extra normalization factor in the denominators of Eqs.~(\\ref{eq:LargestEigenvaluePT}, \\ref{eq:OtherEigenvaluesPT}), the expression for the regularization, although related, takes on a different form than that in Refs.~\\cite{Lakshminarayan16, Tomsovic18c}.\n\n\n\\section{Eigenvalue moments of the reduced density matrix}\n\\label{sec:eigenvalue-moments}\n\nTo fully characterize the entanglement of the evolving state, the Schmidt eigenvalue expression in Eq.~(\\ref{eq:SchmidtEigenvaluelneq1Reg}) is used to compute the leading order of general moments analytically and thereby the HCT entropies, good up to and including $\\mathcal{O}(\\sqrt{\\Lambda})$.\n\n\\subsection{General moments} \\label{subsec:GeneralMomentsCalc}\n\nConsider the ensemble average of the general moments $\\mu_\\alpha$, Eq.~\\eqref{eq:GenericMoments}, of the Schmidt eigenvalues. The largest eigenvalue must be separated out from the others and two integrals considered. First, consider general moments of the sum of all the Schmidt eigenvalues other than the largest, i.e.\n\\begin{equation}\n\\overline{ \\sum_{l \\neq 1} \\lambda_l^\\alpha(t;\\Lambda) } = C_2(\\alpha;t) \\, \\sqrt{\\Lambda} \\,,\n\\end{equation}\nwhere after rescaling $s$ to $z$ in Eq.~(\\ref{eq:R-function})\n\\begin{align}\n& C_2(\\alpha;t) = \\int_{-\\infty}^\\infty \\dd{z} \\int_0^\\infty \\dd{w} \\overline{R(z,w)} \\frac{4^\\alpha w^\\alpha}{(z^2+4w)^\\alpha} \\nonumber \\\\\n& \\qquad \\qquad \\quad \\times \\sin^{2\\alpha}\\bigg( \\frac{t}{2} \\sqrt{z^2+4w}\\,\\bigg) \\nonumber \\\\\n& \\qquad \\qquad = \\int_{-\\infty}^\\infty \\dd{z} \\int_0^\\infty \\dd{w} \\rho(w) \\frac{4^\\alpha w^\\alpha}{(z^2+4w)^\\alpha} \\nonumber \\\\\n& \\qquad \\qquad \\quad \\times \\sin^{2\\alpha}\\bigg( \\frac{t}{2} \\sqrt{z^2+4w}\\,\\bigg). \\label{eq:C2Integral}\n\\end{align}\nThe evaluation of this integral is discussed in the next subsection and\nApp.~\\ref{app:C-2-alpha-t-derivation}.\nNow focusing on the ensemble average of the largest Schmidt eigenvalue,\n\\begin{align}\n& \\overline{\\lambda_1^\\alpha(t;\\Lambda)} = \\overline{\\bigg( 1 - \\sum_{l \\neq 1} \\lambda_l(t;\\Lambda)\\bigg)^\\alpha} \\nonumber \\\\\n& \\qquad\\,\\,\\quad= \\overline{\\Bigg[ 1 - \\int_{-\\infty}^\\infty \\dd{z} \\int_0^\\infty \\dd{w} R(z,w) \\frac{4 w}{z^2+4w}} \\nonumber \\\\\n& \\qquad \\qquad \\quad \\overline{\\times \\sin^{2\\alpha}\\bigg( \\frac{t}{2} \\sqrt{z^2+4w}\\,\\bigg) \\Bigg]^\\alpha}\\,. \\label{eq:LargestEigenvalueExactIntegral}\n\\end{align}\nBefore computing the above for a general power $\\alpha$, consider the $\\alpha =2$ case. Expanding the above expression gives a square of the integral term, which is a quadruple integral containing the product $R(z_1,w_1) R(z_2,w_2)$. Equation (\\ref{threepoint}) has two contributions, the diagonal term, where $(z_1,w_1) = (z_2,w_2)$, and the off-diagonal one. For an uncorrelated spectrum, any multi-point spectral correlation function is unity. Thus\n\\begin{align}\n\\overline{\\lambda_1^2(t;\\Lambda)} = &\\; 1 - 2 C_2(1;t)\\sqrt{\\Lambda} + C_2(2;t) \\sqrt{\\Lambda} \\nonumber \\\\\n& + C^2_2(1;t)\\, \\Lambda,\n\\end{align}\nwhere the off-diagonal term, $R_3(z_1,z_2)$, is responsible for the $\\mathcal{O}(\\Lambda)$ term.\nThis illustrates that to the leading $\\mathcal{O}(\\sqrt{\\Lambda})$, the diagonal term alone suffices, and the other terms contribute to higher than leading order. This simplifies the $\\overline{\\lambda_1^\\alpha}$ computation, where after the binomial expansion of Eq.~(\\ref{eq:LargestEigenvalueExactIntegral}), keeping only the terms contributing to the leading order gives,\n\\begin{align}\n& \\overline{ \\lambda_1^\\alpha(t;\\Lambda) } = 1 + \\sum_{p=1}^\\infty (-1)^p \\binom{\\alpha}{p} \\overline{ \\sum_{l\\neq 1} \\lambda_l^p }.\n\\end{align}\nFinally, the general moments Eq.~\\eqref{eq:GenericMoments}\nof the Schmidt eigenvalues for $\\alpha >1\/2$ are given by\n\\begin{equation}\n\\overline{\\mu_{\\alpha}(t;\\Lambda) } = 1 + \\sum_{p=1}^\\infty (-1)^p \\binom{\\alpha}{p} \\overline{ \\sum_{l\\neq 1} \\lambda_l^p }\\, +\\, \\overline{ \\sum_{l\\neq 1} \\lambda_l^\\alpha } .\n\\end{equation}\nThese can be written as\n\\begin{equation} \\label{eq:mu-alpha-t---C-alpha-t}\n\\overline{ \\mu_{\\alpha}(t;\\Lambda) } = 1 - C(\\alpha;t) \\sqrt{\\Lambda},\n\\end{equation}\nwhere\n\\begin{equation} \\label{eq:C-alpha-t-via-C-2}\nC(\\alpha;t) = \\sum_{p=1}^\\infty (-1)^{p+1} \\binom{\\alpha}{p} C_{2}(p;t) - C_{2}(\\alpha;t).\n\\end{equation}\nThese functions are central for the analytical\ndescription of the entropies and shown\nin Fig.~\\ref{fig:Calpha} for the COE and the CUE.\nA particular feature is the overshooting before the saturation sets in.\nFor the CUE the location of the maxima occurs slightly later in $t$\nthan for the COE and also the saturation regime is reached\nslightly later.\nMoreover, the saturation value is slightly larger than in the COE case.\n\n\\begin{figure}\n\\includegraphics[width=8.6cm]{fig_Calpha_COE.pdf}\n\n\\includegraphics[width=8.6cm]{fig_Calpha_CUE.pdf}\n\n\\caption{Plot of $C(\\alpha;t)$ for $\\alpha=2$ (solid),\n $\\alpha=3$ (dashed), $\\alpha=4$ (dot-dashed),\nand $\\pdv*{C(\\alpha;t)}{\\alpha}|_{\\alpha \\rightarrow 1}$ (dotted)\nfor (a) COE and (b) CUE.}\n\\label{fig:Calpha}\n\\end{figure}\n\n\nIn addition, it can be shown that $\\overline{ \\mu_\\alpha (t;\\Lambda) }$ is evaluated up to and including $\\mathcal{O}(\\sqrt{\\Lambda})$ by the first and second largest Schmidt eigenvalues\n\\begin{equation}\n \\overline{ \\mu_\\alpha (t;\\Lambda) }\n = \\overline{\\lambda_1^\\alpha(t;\\Lambda)} +\n \\overline{\\lambda_2^\\alpha(t;\\Lambda)},\n\\end{equation}\nas other Schmidt eigenvalues ($l > 2$) do not contribute to $ \\mathcal{O} ( \\sqrt{ \\Lambda } ) $.\nThis relation is vital for recursively invoking perturbation theory\nin Sec.~\\ref{sec:NonPerturbative}\nin order to extend the results beyond the perturbative regime.\n\n\\subsection{Entropies}\n\nThe HCT entropies can be computed in the perturbation regime using the results for the average eigenvalue moments. For $\\alpha \\neq 1$ one has\n\\begin{equation}\n\\overline{ S_\\alpha(t;\\Lambda)} = \\frac{C(\\alpha;t)}{\\alpha -1} \\sqrt{\\Lambda},\n\\end{equation}\nwhere $C(\\alpha; t)$ is given by Eq.~\\eqref{eq:C-alpha-t-via-C-2}.\nThis requires the computation of $C_2(\\alpha; t)$,\nwhich is done in App.~\\ref{app:C-2-alpha-t-derivation},\nand leads to\n\\begin{equation} \\label{eq:C-2-alpha-t}\n C_{2}(\\alpha;t) = 2^\\alpha \\sum_{q=0}^\\infty \\sum_{m=0}^q (-1)^q\n \\binom{\\alpha}{q} a_{qm} f_{m}(\\alpha;t),\n\\end{equation}\nwhere\n\\begin{equation}\na_{qm} = \\binom{q}{\\frac{q-m}{2}}\n \\left[ \\frac{1+(-1)^{q-m}}{2^q(1+\\delta_{m,0})} \\right]\n\\end{equation}\nand\n\\begin{align}\n f_m(\\alpha;t) = & \\int_{-\\infty}^\\infty \\dd{z} \\int_0^\\infty\n \\dd{w} \\frac{w^\\alpha \\rho(w)}{(z^2+4w)^\\alpha} \\nonumber \\\\\n & \\qquad \\times \\cos(m t \\sqrt{z^2+4w}\\,).\n\\end{align}\nExplicit expressions for $f_m(\\alpha;t)$\nfor the COE and CUE are derived in App.~\\ref{app:C-2-alpha-t-derivation-COE}\nand \\ref{app:C-2-alpha-t-derivation-CUE}, respectively.\n\n\n\n\\subsection{Discussion}\n\nTo discuss some qualitative features, the case $\\alpha = 2$, which corresponds\nto the linear entropy, is considered here.\nBy Eqs.~(\\ref{eq:GenericHCTEntropy}, \\ref{eq:mu-alpha-t---C-alpha-t})\none has $S_2(t;\\Lambda) = C(2; t) \\sqrt{\\Lambda}$.\nIn case of the COE\n\\begin{align}\n& C(2;t) = 4 \\pi t \\big( \\text{e}^{-t^2} [ \\{ 1 + 2 t^2 \\} I_0(t^2) + 2 t^2 I_1(t^2) ] \\nonumber \\\\\n& \\qquad \\qquad - 4 t^2 \\text{e}^{-4 t^2} [ I_0(4 t^2) + I_1(4 t^2) ] \\, \\big),\n\\end{align}\nwhere $I_n(z)$ is the modified Bessel function of the first kind\n\\cite[Eq.~10.25.2]{DLMF}.\nWhereas for the CUE case\n\\begin{align}\n& C(2;t) = \\pi t \\Big( 3 \\text{e}^{-t^2} - \\frac{1}{2}\\text{e}^{-4 t^2} \\Big) + \\pi^{3\/2} \\text{erf}(t) \\Big( \\frac{1}{2} + 3 t^2 \\Big) \\nonumber \\\\\n& \\qquad \\qquad + \\pi^{3\/2} \\text{erf}(2 t) \\Big( \\frac{1}{8} - 3 t^2 \\Big),\n\\end{align}\nwhere $\\text{erf}(z)$ is the error function.\nFor both COE and CUE cases, $C(2;t)$ for small $t$ has the expansion\n\\begin{align} \\label{eq:C-2-t}\n & C(2;t) = 4 \\pi t + \\mathcal{O}(t^3)\n\\end{align}\nand naturally gives linear-in-time entropy growth for short time $t$.\nThe same is true for other $\\alpha$-entropies, except for $\\alpha = 1$,\nfor which the leading term is of the order $\\mathcal{O}(t\\,\\ln t)$. In fact, it can be shown that for both COE and CUE cases, with $\\alpha > 1$ and short time $t$,\n\\begin{equation}\n\\dv{}{t} C(\\alpha;t) \\big|_{t \\rightarrow 0} = 2 \\pi \\alpha. \\label{eq:InitialRateCalpha}\n\\end{equation}\n\nIn the limit $t \\rightarrow \\infty$, saturation values of the entropies\ncan be obtained from\n\\begin{equation}\n S_2(\\infty; \\Lambda) = \\frac{C(\\alpha; \\infty)}{\\alpha-1} \\sqrt{\\Lambda},\n\\end{equation}\nwhich are of the order $\\mathcal{O}(\\sqrt{\\Lambda}\\,)$.\nUsing the explicit expressions for $f_m(\\alpha;t)$\nderived in App.~\\ref{app:C-2-alpha-t-derivation-COE}, \\ref{app:C-2-alpha-t-derivation-CUE}\none sees that in the limit $t \\rightarrow \\infty$,\n$f_m(\\alpha;t)$ vanish for all $m \\neq 0$.\n Using this fact, an expression for saturation value\nof the $\\alpha$-entropies ($\\alpha \\neq 1$) can be derived as\n\\begin{align}\n\\overline{S_\\alpha(\\infty, \\Lambda)} & = \\Bigg(\\alpha \\,\\, _{3}F_2(1\/2,3\/2,1-\\alpha\\,;\\,2,2\\,;\\,1) \\nonumber \\\\\n&\\; \\qquad \\quad - \\frac{2}{\\pi} \\,\\,\\frac{\\Gamma(\\alpha-1\/2)\\,\\Gamma(\\alpha+1\/2)}{\\Gamma(\\alpha)\\,\\Gamma(\\alpha+1)} \\Bigg) \\nonumber \\\\\n& \\; \\qquad \\times \\frac{\\sqrt{\\Lambda}}{\\alpha-1}\\begin{cases}\n\t\t\t\t\t\t\t\\sqrt{2\\pi} \\; \\quad \\quad \\text{for COE,} \\\\\n\t\t\t\t\t\t\t\\pi^{3\/2}\/2 \\, \\; \\quad \\text{for CUE.}\n\t\t\t\t\t\t \\end{cases}\n \\label{eq:sat-S-alpha}\n\\end{align}\nHere $_{m}F_n$ is a generalized hypergeometric function\n\\cite[Eq.~35.8.1]{DLMF} defined by\n\\begin{align}\n& _{m}F_n ( a_1 ,\\ldots, a_m ; b_1 ,\\ldots, b_n ; z) = \\sum_{k=0}^\\infty \\frac{(a_1)_k \\ldots (a_m)_k}{(b_1)_k \\ldots (b_n)_k} \\frac{z^k}{k!},\n\\end{align}\nwhere $(a)_k = \\Gamma(a+k)\/\\Gamma(a)$ is Pochhammer's symbol.\nEquation~\\eqref{eq:sat-S-alpha} shows that the saturation values\nfor both COE and CUE scale with $\\sqrt{\\Lambda}$\nand that the CUE case leads to a slightly (11\\%) larger value.\n\n\n\nFor the linear entropy, $\\alpha=2$, Eq.~\\eqref{eq:sat-S-alpha} simplifies to\n\\begin{align} \\label{eq:S2-sat-COE-CUE}\n\\overline{S_2(\\infty;\\Lambda)} = \\sqrt{\\Lambda}\n \\begin{cases}\n 5 \\sqrt{\\pi\/8} & \\text{for COE}, \\\\\n 5 \\pi^{3\/2}\/8 & \\text{for CUE}.\n \\end{cases}\n\\end{align}\nFor the special case of the von Neumann entropy for $\\alpha=1$,\n$\\lim_{t\\rightarrow \\infty} \\pdv*{C(\\alpha;t)}{\\alpha}|_{\\alpha \\rightarrow 1}$\nneeds to be computed. It can be shown that\n\\begin{align} \\label{eq:saturation-perturbatively}\n\\overline{S_1(\\infty;\\Lambda)}\n= & \\Big(\n 4 \\ln 2- \\frac{3}{16}\\, _{4}F_3(1,1,3\/2,5\/2 \\,;\\,2,3,3\\,;\\,1) \\Big)\n\\nonumber \\\\\n& \\times \\sqrt{\\Lambda} \\begin{cases}\n\t\t\t\t\t\t\t\\sqrt{2\\pi} \\; \\quad \\quad \\text{for COE,} \\\\\n\t\t\t\t\t\t\t\\pi^{3\/2}\/2 \\, \\; \\quad \\text{for CUE.}\n\t\t\t\t\t\t \\end{cases}\n\\end{align}\nAn extension to the non-perturbative result will be discussed\nin Sec.~\\ref{sec:long-time-saturation}.\n\nIn the perturbative regime, if the entropies $\\overline{ S_\\alpha (t;\\Lambda)}$ are scaled with respect to\ntheir saturation values,\n\\begin{equation}\n\\overline{ \\mathcal{S}_\\alpha (t) }\n = \\frac{\\overline{S_\\alpha(t; \\Lambda)}}\n {\\overline{S_\\alpha(\\infty; \\Lambda)}},\n\\label{eq:Universal-Scaling}\n\\end{equation}\nthey do not depend on the transition parameter,\nleading to one universal curve for each $\\alpha$ described\nby the prediction\n\\begin{equation}\n\\overline{ \\mathcal{S}_\\alpha (t) } = \\frac{C(\\alpha;t)}{C(\\alpha;\\infty)}.\n \\label{eq:UniversalCurveTheory}\n\\end{equation}\nThis universal property is depicted for the linear entropy in Fig.~\\ref{fig:UniversalCurvePlot} for various $\\Lambda$-values. As $\\Lambda$ goes beyond the perturbation regime, departure from the universal curve is seen due to the breakdown of the perturbation theory. In the forthcoming section, the extension of the theory to the non-perturbative regime is discussed.\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=8.6cm]{fig_CUE_universal.pdf}\n\n \\caption{\\label{fig:UniversalCurvePlot} Scaled linear entropy\n $\\overline{ \\mathcal{S}_2(t) }$, Eq.~\\eqref{eq:Universal-Scaling}, for the\n random matrix transition ensemble in Eq.~(\\ref{eq:GenericFloquetRMT}) for\n the CUE case for $\\Lambda = 10^{-6}$ (magenta circles),\n $\\Lambda = 10^{-4}$ (red triangles), $\\Lambda = 10^{-2}$ (blue squares),\n and $\\Lambda = 1$ (green diamonds); The theoretical prediction\n Eq.~(\\ref{eq:UniversalCurveTheory}) for $\\alpha =2$ is shown as solid curve.}\n\\end{figure}\n\n\n\\section{Non-perturbative regime} \\label{sec:NonPerturbative}\n\nThe results obtained from the perturbation theory can be extended to the\nnon-perturbative regime to produce the full transition and the saturation\nvalues by employing the recursively embedded perturbation theory technique as\ndone in Ref.~\\cite{Lakshminarayan16,Tomsovic18c} for the eigenstates.\n\n\n\\subsection{Full transition}\n\n\nFor small enough $\\Lambda$, the time evolved state $\\ket{jk(n;\\epsilon)} \\equiv \\ket{jk(t;\\Lambda)}$ can be Schmidt decomposed as\n\\begin{equation}\n\\ket{jk(t;\\Lambda)} = \\sqrt{\\lambda_1(t;\\Lambda)} \\,\\ket{(jk)_1} + \\sqrt{\\lambda_2(t;\\Lambda)} \\,\\ket{(jk)_2},\n\\end{equation}\nsuch that $\\lambda_1 + \\lambda_2 = 1$, where the time-dependent phase-factor is absorbed into the definition of the Schmidt eigenvectors $\\ket{ (jk)_l }$. Now increasing the interaction strength, another unperturbed state energetically close to $\\ket{(jk)_1}$ will contribute to $\\ket{jk(n;\\epsilon)}$,\n\\begin{align}\n& \\ket{jk(t;\\Lambda)} = \\sqrt{\\lambda_1'(t;\\Lambda)} \\,\\bigg(\\sqrt{\\lambda_1(t;\\Lambda)} \\,\\ket{(jk)_1} + \\nonumber \\\\\n& \\qquad \\qquad \\quad \\sqrt{\\lambda_2(t;\\Lambda)} \\,\\ket{(jk)_2}\\bigg) + \\sqrt{\\lambda_2'(t;\\Lambda)} \\, \\ket{(jk)_3},\n\\end{align}\nwhere $\\lambda_{1,2}'$ follow same statistical properties as the unprimed ones. Thus the purity is $\\mu_2' = \\lambda_1^{'2} \\lambda_1^2 + \\lambda_1^{'2} \\lambda_2^2 + \\lambda_2^{'2}$ giving\n\\begin{align}\n\\mu_2' - \\mu_2 = -(1-\\lambda_1^{'2} - \\lambda_2^{'2})\\mu_2 + \\lambda_2^{'2} (1-\\mu_2).\n\\end{align}\n\n\n\\begin{figure}\n\\includegraphics[width=8.4cm]{fig_COE_TP_1e-06_N_50.pdf}\n\n\\includegraphics[width=8.4cm]{fig_COE_TP_0p0001_N_50}\n\n\\includegraphics[width=8.4cm]{fig_COE_TP_0p01_N_50.pdf}\n\n\\includegraphics[width=8.4cm]{fig_COE_TP_1_N_50}\n\\caption{\\label{fig:COE_Salpha} Entropies $\\overline{ S_\\alpha }$ for the COE\n case with $N_A=N_B=50$ for (a) $\\Lambda=10^{-6}$, (b) $\\Lambda=10^{-4}$, (c)\n $\\Lambda=10^{-2}$, and (d) $\\Lambda=1$ for $\\alpha=1$ (green diamonds),\n $\\alpha=2$ (magenta circles), $\\alpha=3$ (red triangles), and $\\alpha=4$\n (blue squares). Black lines show the corresponding theory curves,\n Eq.~(\\ref{eq:alphaEntropyTheory}).}\n\\end{figure}\n\n\\begin{figure}\n\\includegraphics[width=8.4cm]{fig_CUE_TP_1e-06_N_50.pdf}\n\n\\includegraphics[width=8.4cm]{fig_CUE_TP_0p0001_N_50}\n\n\\includegraphics[width=8.4cm]{fig_CUE_TP_0p01_N_50.pdf}\n\n\\includegraphics[width=8.4cm]{fig_CUE_TP_1_N_50}\n\\caption{\\label{fig:CUE_Salpha} Entropies $\\overline{ S_\\alpha }$ for the CUE\n case with $N_A=N_B=50$ for (a) $\\Lambda=10^{-6}$, (b) $\\Lambda=10^{-4}$, (c)\n $\\Lambda=10^{-2}$, and (d) $\\Lambda=1$ for $\\alpha=1$ (green diamonds),\n $\\alpha=2$ (magenta circles), $\\alpha=3$ (red triangles), and $\\alpha=4$\n (blue squares). Black lines show the corresponding theory curves,\n Eq.~(\\ref{eq:alphaEntropyTheory}).}\n\\end{figure}\n\nFor a given $\\alpha$, following this technique and replacing $\\lambda_{1,2}^{'\\alpha}$ with their ensemble average, a differential equation for the moments $\\overline{ \\mu_\\alpha (t;\\Lambda) }$ can be derived, good up to $\\mathcal{O}(\\sqrt{\\Lambda})$,\n\\begin{equation}\n\\pdv{\\overline{ \\mu_\\alpha(t;\\Lambda) }}{\\sqrt{\\Lambda}} = - C(\\alpha;t) \\overline{ \\mu_\\alpha (t;\\Lambda)}.\n\\end{equation}\nThis has a solution of the form (valid for the infinite dimensional case)\n\\begin{equation}\n\\overline{ \\mu_\\alpha(t;\\Lambda)} \\approx \\exp(-C(\\alpha;t) \\sqrt{\\Lambda}).\n\\end{equation}\nIn the limit $\\Lambda \\rightarrow \\infty$, and for large (but finite)\ndimensionality $N=N_A =N_B$, the moments tend to the random matrix result\n\\begin{equation}\n\\overline{ \\mu_\\alpha^\\infty } = \\mathcal{C}_\\alpha\/N^{\\alpha-1},\n\\end{equation}\nwhere the $\\mathcal{C}_\\alpha$ are Catalan numbers \\cite[\\S26.5.]{DLMF}.\nFor the $N_A \\neq N_B$ case such an expression can be found following Ref.~\\cite{Sommers04}. Incorporating this limit\ninto the above differential equation solution gives an approximate expression for the moments valid for any $\\Lambda$,\n\\begin{equation}\n\\overline{ \\mu_\\alpha (t;\\Lambda) } \\approx \\exp(- \\frac{C(\\alpha;t)}{1-\\overline{\\mu_\\alpha^\\infty}}\\sqrt{\\Lambda})(1-\\overline{\\mu_\\alpha^\\infty}) + \\overline{ \\mu_\\alpha^\\infty}.\n\\end{equation}\nUsing the definition of the HCT entropies (\\ref{eq:GenericHCTEntropy})\ngives\n\\begin{equation}\n\\overline{S_\\alpha(t;\\Lambda) } \\approx \\bigg[ 1 - \\exp(-\\frac{C(\\alpha;t)}{(\\alpha-1)\\overline{ S_\\alpha^\\infty}}\\sqrt{\\Lambda})\\bigg] \\overline{S_\\alpha^\\infty}, \\label{eq:alphaEntropyTheory}\n\\end{equation}\nwhere\n\\begin{equation}\n\\overline{ S_\\alpha^\\infty} = \\frac{1-\\mathcal{C}_\\alpha N^{1-\\alpha}}{\\alpha-1}.\n\\end{equation}\nTo apply Eq.~\\eqref{eq:alphaEntropyTheory}\none has to use for $C(\\alpha; t)$ the results\ncorresponding to the CUE or the COE,\nas given by Eq.~\\eqref{eq:C-alpha-t-via-C-2}.\nWhen $\\Lambda$ is large, however,\nthere is no difference between CUE and COE\ndue to the same scaling of $C(\\alpha; t)$,\nas for example in Eq.~\\eqref{eq:C-2-t} for $\\alpha=2$.\n\n\nThe result Eq.~(\\ref{eq:alphaEntropyTheory}) is in agreement with\nnumerical computations for both the COE, see Fig.~\\ref{fig:COE_Salpha},\nand the CUE, see Fig.~\\ref{fig:CUE_Salpha}.\nFor these numerical calculations, 20 realizations of the random matrix model Eq.~\\eqref{eq:GenericFloquetRMT} for $N_A=N_B=50$ have been used, leading to a total of $5 \\times 10^4$ initially unentangled eigenstates $\\ket{jk}$ used for averaging. This amount of averaging is\nparticularly relevant for small values of $\\Lambda$\nfor which the time evolution of the entanglement of the individual\nstates shows strong fluctuations from one state to another.\nThese are also the origin of the small fluctuations\nseen in both figures for $\\Lambda=10^{-6}$\nfor the von Neumann entropy $\\overline{S_1(t; \\Lambda)}$,\nwhich is the most sensitive of the considered entropies.\nMoreover, at small $\\Lambda$, finite $N$ effects\nbecome visible, in particular for the COE,\ndue to the small overall amount of entanglement.\nIncreasing the matrix dimension of the subsystems\nto $N=100$ improves the agreement with the theoretical prediction\n(not shown).\nFor $\\Lambda=10^{-4}$ and $\\Lambda=10^{-2}$ excellent\nagreement of the numerically computed entropies and\nthe theory is found.\nFor $\\Lambda=1$ again the von Neumann entropy shows\nsmall deviations from the theoretical prediction.\n\n\\subsection{Long-time saturation}\n\\label{sec:long-time-saturation}\n\n\\begin{figure}[b]\n\\centering\n \\includegraphics[width=8.4cm]{fig_saturation_logscale.pdf}\n\\caption{\\label{fig:saturation}\nSaturation values of the linear entropy, $\\overline{S_2(\\infty;\\Lambda)}$, as a function of $\\Lambda$\nfor the COE (blue squares) and CUE (red diamonds) in comparison\nwith the prediction Eq.~\\eqref{eq:sat-alphaEntropyTheory}\n(dashed and solid black lines representing COE and CUE, respectively).}\n\\end{figure}\n\n\nUsing the result Eq.~\\eqref{eq:alphaEntropyTheory}\none can also perform the long-time limit to\nobtain a prediction for the saturation\nvalues $\\overline{S_\\alpha(\\infty;\\Lambda) }$\ngoing beyond the perturbative result Eq.~\\eqref{eq:saturation-perturbatively}.\nThus one gets\n\\begin{equation}\n \\overline{S_\\alpha(\\infty;\\Lambda) } = \\bigg[ 1 - \\exp(-\\frac{C(\\alpha;\\infty)}{(\\alpha-1)\\overline{ S_\\alpha^\\infty}}\\sqrt{\\Lambda})\\bigg] \\overline{S_\\alpha^\\infty}. \\label{eq:sat-alphaEntropyTheory}\n\\end{equation}\nFor large $\\Lambda$ the exponential becomes very small\nso that the saturation reaches $\\overline{S_\\alpha^\\infty}$.\nHowever, for small $\\Lambda$ a reduced saturation value is obtained.\nFigure~\\ref{fig:saturation} illustrates\nthis for the linear entropy for the COE and CUE\nwhere Eq.~\\eqref{eq:S2-sat-COE-CUE} is used for $C(\\alpha, \\infty)$.\nVery good agreement of the prediction with the numerical results\nis found.\nUp to $\\Lambda=10^{-1}$ the saturation value\nfollows Eq.~\\eqref{eq:S2-sat-COE-CUE}\nand then the behavior given by Eq.~\\eqref{eq:sat-alphaEntropyTheory}\nsets in.\nThe saturation values for the COE are below\nthose of the CUE but eventually both approach $\\overline{S_2^\\infty}$.\n\n\\section{Coupled kicked rotors}\n\\label{sec:coupled-kicked-rotors}\n\n\\begin{figure}\n\\includegraphics[width=8.4cm]{fig_KR_TP_1e-06_N_100.pdf}\n\n\\includegraphics[width=8.4cm]{fig_KR_TP_0p0001_N_100.pdf}\n\n\\includegraphics[width=8.4cm]{fig_KR_TP_0p01_N_100.pdf}\n\n\\includegraphics[width=8.4cm]{fig_KR_TP_1_N_100.pdf}\n\n\\caption{\\label{fig:KR_Salpha} Entropies $\\overline{ S_\\alpha }$ for the coupled kicked rotors with completely broken time-reversal invariance and $N=100$, for (a) $\\Lambda=10^{-6}$, (b) $\\Lambda=10^{-4}$, (c)\n $\\Lambda=10^{-2}$, and (d) $\\Lambda=1$ for $\\alpha=1$ (green diamonds),\n $\\alpha=2$ (magenta circles), $\\alpha=3$ (red triangles), and $\\alpha=4$\n (blue squares). Black lines show the corresponding theory curves,\n Eq.~(\\ref{eq:alphaEntropyTheory}).}\n\\end{figure}\n\nA bipartite system whose subsystems exhibit classical chaotic motion is considered here to compare against the universal entanglement dynamics results derived from random matrix theory. The knowledge of $\\Lambda$ and its relation to the system dependent details is crucial for the comparison.\nIn case of a system whose subsystems are kicked rotors quantized on the unit torus, it is possible to analytically find $\\Lambda$ as a function of system dependent details as shown in Refs.~\\cite{Srivastava16,Tomsovic18c}. The Floquet unitary operator of the system has the form given by Eq.~(\\ref{eq:GenericFloquet}) where the subsystem Floquet operator for one kicked rotor is\n\\begin{equation}\nU_A = \\exp[- i p_A^2\/(2\\hbar)] \\exp(-i V_A\/\\hbar),\n\\end{equation}\nwith kicking potential given by\n\\begin{equation}\nV_A = K_A \\cos(2\\pi q_A)\/4\\pi^2,\n\\end{equation}\nwhere $K_A$ is the kicking strength. Similarly for subsystem $B$. The entangling operator is\n\\begin{equation}\nU_{AB}(b) = \\exp(-i b V_{AB}\/\\hbar),\n\\end{equation}\nwhere the interaction potential is\n\\begin{equation}\nV_{AB} = \\frac{1}{4\\pi^2} \\cos[2\\pi(q_A+q_B)].\n\\end{equation}\nThe angle variables $q_j$ is restricted to the interval $[0,1)$, and similarly for the momenta $p_j$. This restriction leads to a 4-dimensional torus phase space for the corresponding classical system \\cite{Froeschle71,Lakshminarayan2001,Richter14}. The kicking strengths $(K_A,\\, K_B) = (10,14),\\,(18,22),\\, \\ldots$ with up to 20 realizations are chosen such that the classical dynamics is chaotic. The boundary conditions are chosen such that both time-reversal invariance and parity symmetry are broken. Thus the subsystem spectral fluctuations are approximately like those of the CUE. In addition we use $N=N_A = N_B = 100$ for the numerical computations. The transition parameter for the coupled kicked rotors is \\cite{Srivastava16,Tomsovic18c,HerKieFriBae2019:p}\n\\begin{equation}\n\\Lambda_{\\text{KR}} \\simeq \\frac{N^2}{4\\pi^2} \\left(1-J_0^2(Nb\/2\\pi) \\right) \\approx \\frac{N^4 b^2}{32 \\pi^4},\n\\end{equation}\nwhere $J_0(\\cdot)$ is the Bessel function of first kind\n\\cite[Eq.~10.2.2]{DLMF},\nand the approximation is true when $Nb \\ll 1$. In Fig.~\\ref{fig:KR_Salpha}, the entanglement dynamics for various $\\Lambda$-values of the coupled kicked rotors is shown against the theory given by Eq.~(\\ref{eq:alphaEntropyTheory}).\nOverall good agreement is found\nwith some small deviations for the von Neumann entropy\nwhich are similar to those found for the\nCUE case shown in Fig.~\\ref{fig:CUE_Salpha}.\n\n\\section{Summary and outlook}\n\\label{sec:summary-and-outlook}\n\nAn analytic theory is given in this paper for the rate of entanglement production for a quenched system as a function of the interaction strength between chaotic subsystems. In particular, all the expressions are given in terms of the universal transition parameter $\\Lambda$. It is shown that in the perturbative regime for an initial product of subsystem eigenstates (a so-called quench), the entanglement saturates at very small values proportional to $\\sqrt{\\Lambda}$. Furthermore, in the same regime, once the appropriate time scale is properly identified and the entanglement entropies are scaled by their saturation value, there exists a single universal entropy production curve: for a given system size, the interaction strength determines $\\Lambda$, which determines the time scale and saturation values, and there is no other dependence in the entropy production beyond that. The universal curve has an overshoot, which is slightly more pronounced for the time reversal non-invariant case, and then it settles down to a saturation value. As $\\Lambda$ increases, the perturbation regime eventually breaks down, roughly for $\\Lambda \\gtrsim 10^{-2}$, as illustrated in Fig.~\\ref{fig:UniversalCurvePlot}.\n\nAs for the full eigenstates of the interacting system \\cite{Tomsovic18c}, it was also possible here to recursively embed the perturbation theory. This enables a description of the full transition in entropy production behaviors as a function of subsystem interaction strength and size, the limiting behaviors being no entanglement entropy production for non-interacting systems, and for strongly interacting systems the production behavior seen for initially random product states. The expressions are uniformly valid for all times and interaction strengths. It also turns out that the initial entropy production rate is even independent of whether time reversal symmetry is preserved or not.\n\nThe present study also raises various interesting\nquestions to be addressed in the future:\nthe considered case of initial states given by direct products of subsystem eigenstates has the crucial property\nthat the automatic Schmidt decomposition holds,\nwhich allows for a perturbative treatment.\nIf one considers instead,\nfor example, sums of such eigenstates or\ndirect products of subsystem random vectors, then a much faster entanglement\ngeneration occurs, which requires a completely different\ntheoretical description.\nMoreover, although not shown, the fluctuations of the entropies seen from one initial\nstate to another depend dramatically on whether it is subsystem eigenvectors or random states which are being considered. This should also be reflected in the statistics\nof Schmidt eigenvalues, which are expected to show\nheavy-tailed distributions as was found before in the case of eigenstates.\nAnother interesting ensemble for the case of a dynamical system\nlike the coupled kicked rotors are coherent states\nas initial states. There one will have an initial phase\nfor which the entanglement only grows very slowly\nup to the Ehrenfest time beyond which a fast increase\nof entanglement occurs.\nFinally, bipartite many-body systems, like an interacting spin-chain,\nshould share many of the features\nof the entanglement production demonstrated here,\nand at the same time also allow for even more possibilities\nof initial states.\n\n\n\\acknowledgments\n\nWe would like to thank Maximilian Kieler for useful discussions.\nOne of the authors (ST) gratefully acknowledges support for visits\nto the Max-Planck-Institut f\\\"ur Physik komplexer Systeme.\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section*{Figure Captions\\markboth\n {FIGURECAPTIONS}{FIGURECAPTIONS}}\\list\n {Figure \\arabic{enumi}:\\hfill}{\\settowidth\\labelwidth{Figure\n999:}\n \\leftmargin\\labelwidth\n \\advance\\leftmargin\\labelsep\\usecounter{enumi}}}\n\\let\\endfigcap\\endlist \\relax\n\\def\\tablecap{\\section*{Table Captions\\markboth\n {TABLECAPTIONS}{TABLECAPTIONS}}\\list\n {Table \\arabic{enumi}:\\hfill}{\\settowidth\\labelwidth{Table\n999:}\n \\leftmargin\\labelwidth\n \\advance\\leftmargin\\labelsep\\usecounter{enumi}}}\n\\let\\endtablecap\\endlist \\relax\n\\def\\reflist{\\section*{References\\markboth\n {REFLIST}{REFLIST}}\\list\n {[\\arabic{enumi}]\\hfill}{\\settowidth\\labelwidth{[999]}\n \\leftmargin\\labelwidth\n \\advance\\leftmargin\\labelsep\\usecounter{enumi}}}\n\\let\\endreflist\\endlist \\relax\n\\def\\list{}{\\rightmargin\\leftmargin}\\item[]{\\list{}{\\rightmargin\\leftmargin}\\item[]}\n\\let\\endquote=\\endlist\n\\makeatletter\n\\newcounter{pubctr}\n\\def\\@ifnextchar[{\\@publist}{\\@@publist}{\\@ifnextchar[{\\@publist}{\\@@publist}}\n\\def\\@publist[#1]{\\list\n {[\\arabic{pubctr}]\\hfill}{\\settowidth\\labelwidth{[999]}\n \\leftmargin\\labelwidth\n \\advance\\leftmargin\\labelsep\n \\@nmbrlisttrue\\def\\@listctr{pubctr}\n \\setcounter{pubctr}{#1}\\addtocounter{pubctr}{-1}}}\n\\def\\@@publist{\\list\n {[\\arabic{pubctr}]\\hfill}{\\settowidth\\labelwidth{[999]}\n \\leftmargin\\labelwidth\n \\advance\\leftmargin\\labelsep\n \\@nmbrlisttrue\\def\\@listctr{pubctr}}}\n\\let\\endpublist\\endlist \\relax\n\\makeatother\n\\newskip\\humongous \\humongous=0pt plus 1000pt minus 1000pt\n\\def\\mathsurround=0pt{\\mathsurround=0pt}\n\\def\\eqalign#1{\\,\\vcenter{\\openup1\\jot \\mathsurround=0pt\n \\ialign{\\strut \\hfil$\\displaystyle{##}$&$\n \\displaystyle{{}##}$\\hfil\\crcr#1\\crcr}}\\,}\n\\newif\\ifdtup\n\\def\\panorama{\\global\\dtuptrue \\openup1\\jot \\mathsurround=0pt\n \\everycr{\\noalign{\\ifdtup \\global\\dtupfalse\n \\vskip-\\lineskiplimit \\vskip\\normallineskiplimit\n \\else \\penalty\\interdisplaylinepenalty \\fi}}}\n\\def\\eqalignno#1{\\panorama \\tabskip=\\humongous\n \\halign to\\displaywidth{\\hfil$\\displaystyle{##}$\n \\tabskip=0pt&$\\displaystyle{{}##}$\\hfil\n \\tabskip=\\humongous&\\llap{$##$}\\tabskip=0pt\n \\crcr#1\\crcr}}\n\\relax\n\n\n\n\\def\\begin{equation}{\\begin{equation}}\n\\def\\end{equation}{\\end{equation}}\n\\def\\begin{eqnarray}{\\begin{eqnarray}}\n\\def\\end{eqnarray}{\\end{eqnarray}}\n\\def\\bar{\\partial}{\\bar{\\partial}}\n\\def\\bar{J}{\\bar{J}}\n\\def\\partial{\\partial}\n\\def\\noindent{\\noindent}\n\n\n\n\\def\\kappa{\\kappa}\n\\def\\rho{\\rho}\n\\def\\alpha{\\alpha}\n\\def\\Alpha{\\Alpha}\n\\def\\beta{\\beta}\n\\def\\Beta{\\Beta}\n\\def\\gamma{\\gamma}\n\\def\\Gamma{\\Gamma}\n\\def\\delta{\\delta}\n\\def\\Delta{\\Delta}\n\\def\\epsilon{\\epsilon}\n\\def\\Epsilon{\\Epsilon}\n\\def\\p{\\pi} \n\\def\\Pi{\\Pi}\n\\def\\chi{\\chi}\n\\def\\Chi{\\Chi}\n\\def\\theta{\\theta}\n\\def\\Theta{\\Theta}\n\\def\\mu{\\mu}\n\\def\\nu{\\nu}\n\\def\\omega{\\omega}\n\\def\\Omega{\\Omega}\n\\def\\lambda{\\lambda}\n\\def\\Lambda{\\Lambda}\n\\def\\s{\\sigma} \n\\def\\Sigma{\\Sigma}\n\\def\\varphi{\\varphi}\n\n\n\\def\\relax{\\rm I\\kern-.18em R}{\\relax{\\rm I\\kern-.18em R}}\n\\def\\relax{\\rm 1\\kern-.35em1}{\\relax{\\rm 1\\kern-.35em1}}\n\n\\renewcommand{\\thesection.\\arabic{equation}}}{\\thesection.\\arabic{equation}}\n\\csname @addtoreset\\endcsname{equation}{section}\n\n\n\n\n\n\n\\def${\\cal N}=4${${\\cal N}=4$}\n\\def\\alpha _S{\\alpha _S}\n\\def\\boldsymbol {\\boldsymbol }\n\n\\newcommand{\\Feyn}[1]{#1\\kern-0.45em\/}\n\n\\headheight 10 pt\n\n\n\n\n\\begin{document}\n\n\\title{\\Large \\bf Dual conformal invariance in the Regge limit}\n\\author{\\large C{\\'e}sar~G{\\'o}mez, Johan~Gunnesson, Agust{\\'i}n~Sabio~Vera \\\\\n{\\it Instituto de F{\\' i}sica Te{\\' o}rica UAM\/CSIC,}\\\\ \n{\\it Universidad Aut{\\' o}noma de Madrid, E-28049 Madrid, Spain}}\n\n\\maketitle\n\n\\vspace{-9cm}\n\\begin{flushright}\n{\\small IFT--UAM\/CSIC--09--38}\n\\end{flushright}\n\n\\vspace{7cm}\n\\begin{abstract}\n\\noindent\n\nA dual conformal symmetry, analogous to the dual conformal symmetry observed for the scattering amplitudes of ${\\cal N}=4$ Super Yang-Mills theory, is identified in the Regge limit of QCD. Combined with the original two-dimensional conformal symmetry of the theory, this dual symmetry can potentially explain the integrability of the BFKL Hamiltonian. We also give evidence that the symmetry survives when a subset of unitarity corrections are taken into account by studying briefly the non-planar $2$ to $m$ reggeon transition vertices.\n \n\\end{abstract}\n\n\n\\section{Introduction}\n\nIn the last few years there has been a great deal of progress in the study of gluon scattering amplitudes in \nthe maximally supersymmetric gauge theory in four dimensions, ${\\cal N}=4$ Super Yang-Mills (SYM). One of the most surprising \ndevelopments has been the discovery of a hidden symmetry in the planar ($N_c \\rightarrow \\infty$) limit, coined \nas ``dual super-conformal symmetry''~\\cite{dualconformal, korchemskyconf,confward}, different from the original \nsuper-conformal symmetry of the Lagrangian. This symmetry was uncovered by introducing a new set of variables \n$x_i$, related to the external (all taken as incoming) gluon momenta $p_i$, $i=1\\ldots n$, through\n\\begin{equation}\nx_i - x_{i+1}=p_i \\ , \\label{eq:xi}\n\\end{equation}\nand acts on the $x_i$ just as a four-dimensional conformal symmetry acts on spatial coordinates. The presence \nof this dual symmetry can be understood through the AdS\/CFT correspondence~\\cite{AdSCFT} since it was \nshown~\\cite{fermionicT} that the problem of calculating a given scattering amplitude can be mapped, \nthrough a fermionic T-duality, to that of calculating a light-like Wilson loop with corners at coordinates \ngiven by the $x_i$. This fermionic T-duality maps the string $\\sigma$-model to itself, and the dual conformal \nsymmetry becomes the ordinary symmetry of the space in which the Wilson loop lives. \n\nLike the ordinary conformal symmetry, the dual symmetry is broken by infrared divergences, arising as cusp \ndivergences in the language of Wilson loops. However, the cusp divergences are known to exponentiate, which \nallows the use of the broken symmetry to impose powerful constraints on the amplitudes in the form of anomalous \nWard identities. These identities fix the 4 and 5 point amplitudes completely while the undetermined parts of \nhigher-point amplitudes can only depend on dual-conformal invariants. Also, taken together, the original and \ndual conformal symmetries generate an infinite-dimensional Yangian symmetry~\\cite{plefka}, ordinarily characteristic \nof exactly solvable models.\n\nInteresting properties of ${\\cal N}=4$ amplitudes appear also in their high energy (Regge) limit. In a nutshell, Regge theory \nestablishes the structure of scattering amplitudes when the momentum transfer is small compared to the total \ncenter-of-mass energy. It turns out that ${\\cal N}=4$ amplitudes exhibit Regge-like behaviour at all orders in the 't \nHooft coupling even outside of the Regge limit. In fact, the 4 and 5 point amplitudes are Regge \nexact~\\cite{korchemskyconf, AgustinBartelsLipatov}, \nmeaning that they can always be written in a factorized form characteristic of high energies, irrespective of \nthe values of the kinematical invariants\\footnote{Also, a proposal for the undetermined part of the 6 point amplitude, having the correct Regge behaviour and given in terms of conformal cross-ratios, is given in the latest version of \\cite{AgustinBartelsLipatov2}. }. Furthermore, in the Leading Logarithmic Approximation (LLA) of gluon \namplitudes the Regge limit is independent of the gauge theory, so ${\\cal N}=4$ can give insight into the high energy \nbehaviour of QCD.\n\nIn the Regge limit amplitudes are dominated by the $t$-channel exchange of reggeized gluons (a reggeized gluon is \na collective state of ordinary gluons projected on a colour octet). The bound state of two reggeized gluons when \nprojected on a colour singlet in the $t$-channel is known as the hard (or perturbative) pomeron. The interaction between \nreggeized gluons is governed by the Schr{\\\"o}dinger-like BFKL integral equation~\\cite{BFKL} where the invariant mass of \n$s$-channel gluons can be interpreted as the time variable and its kernel as an effective Hamiltonian living on the \ntwo-dimensional transverse space. This Hamiltonian is free from infrared singularities and carries a yet to be understood \nintegrability~\\cite{BFKLintegrable} \\footnote{Integrability also appears when the gluon composite states are projected onto the\nadjoint representation~\\cite{Lipatov:2009nt}.}.\n\nThe question that arises is if the integrable structures present in ${\\cal N}=4$ SYM can shed some light on the integrability \nfound in the Regge limit. In this region the dynamics of the theory is reduced to the transverse plane, where a two-dimensional \nconformal symmetry was found in the effective Hamiltonian~\\cite{Lipatov1986}. Given the emergence of the Yangian in the \nfour-dimensional case, and that one can heuristically interpret this $SL(2,C)$ as a reduction of the four-dimensional ordinary conformal \nsymmetry, it would then seem natural to look for a dual $SL(2,C)$ symmetry in the high-energy limit\\footnote{Hints in this direction have already appeared in the literature. A similar dual symmetry was exploited in \\cite{Lipatov:2009nt} in order to map supersymmetric multiparticle amplitudes in multiregge kinematics to an integrable open spin chain. In fact, the octet kernel, after subtraction of infrared divergences, can be written in a form manifestly invariant under this symmetry. Also, the BFKL Hamiltonian has holomorphic separability into two pieces which can be written such that they are invariant under the duality transformation $p_i \\to \\rho_i - \\rho_{i+1}\n\\to p_{i+1}$, similar to the change of variables \\eqref{eq:xi}, with the $\\rho$ being the gluon transverse coordinates in\ncomplex notation~\\cite{Lipatov:1998as}.}. It is this question that we address in this Letter, showing that BFKL indeed exhibits covariance under such a dual symmetry.\n\nWe will also study a set of corrections to BFKL, in the form of $2\\rightarrow m$ reggeized gluon transition \nvertices, which also turn out to be dual $SL(2,C)$-covariant. This result is important for high-energy QCD, since the inclusion of such vertices is necessary, at sufficiently high energies, to fulfill unitarity in all channels. \n\n\n\\section{The dual $SL(2,C)$ symmetry}\n\n\n\\begin{figure}[ht]\n\\psfrag{+}{$+$} \\psfrag{=}{$=$} \\psfrag{ka}{$k_A$} \\psfrag{kb}{$k_B$}\n\\psfrag{kp}{$k'$} \\psfrag{F}{$F$} \\psfrag{kamq}{$k_A - q$}\n\\psfrag{kbmq}{$k_B - q$} \\psfrag{x1}{$x_1$} \\psfrag{x2}{$x_2$}\n\\psfrag{x3}{$x_3$} \\psfrag{x4}{$x_4$}\n\\begin{center}\n\\includegraphics{BFKL2.eps}\n\\caption{\\small The BFKL integral equation for the four-point reggeized gluon\nGreen function.} \\label{fig:BFKL}\n\\end{center}\n\\end{figure}\n\nThe scattering amplitude for the 2 to 2 reggeized gluons process in the Regge limit has an iterative structure \ndominated by the exchange in the $t$-channel of a colour singlet. This implies that in the LLA the corresponding \n4 point gluon Green function can be written as the solution to an integral \nequation, the BFKL equation \\footnote{For an introductory treatment of the BFKL equation, see \\cite{rossforshaw}}, shown in Fig.~\\ref{fig:BFKL}. Written in terms of $\\omega$, the Mellin conjugate variable of the \ncenter-of-mass energy (which can be translated into \nthe rapidity, $Y$, of the emitted particles in the $s$-channel)), and the incoming two dimensional momenta it reads \n\\begin{equation}\n\\omega F(\\omega , \\, \\boldsymbol {k}_A,\\, \\boldsymbol {k}_B,\\, \\boldsymbol {q}) = \\delta^{(2)}(\\boldsymbol {k}_A - \\boldsymbol {k}_B) + \\int d^2\\boldsymbol {k}' K(\\boldsymbol {k}_A,\\,\n \\boldsymbol {k}_A-\\boldsymbol {q};\\, \\boldsymbol {k}',\\, \\boldsymbol {k}'-\\boldsymbol {q}) F(\\omega , \\, \\boldsymbol {k}',\\, \\boldsymbol {k}_B ,\\, \\boldsymbol {q}) \\, \\label{eq:BFKLconK} \n\\end{equation}\nwhere the kernel $K(\\boldsymbol {k}_A,\\, \\boldsymbol {k}_A-\\boldsymbol {q};\\, \\boldsymbol {k}',\\, \\boldsymbol {k}'-\\boldsymbol {q})$ is given by\n\\begin{align}\n\\frac{K_R (\\boldsymbol {k}_A,\\, \\boldsymbol {k}_A-\\boldsymbol {q}; -\\boldsymbol {k}'+\\boldsymbol {q}, \\, -\\boldsymbol {k}')}{8\\pi^3\\boldsymbol {k}_A^2(\\boldsymbol {k}' - \\boldsymbol {q})^2} \n+ \\left[ \\omega (\\boldsymbol {k}_A^2) + \\omega ((\\boldsymbol {k}_A-\\boldsymbol {q})^2) \\right]\\delta ^{(2)}(\\boldsymbol {k}_A-\\boldsymbol {k}') \\ . \\label{eq:Kernel}\n\\end{align}\nThe ``real emission'' part\\footnote{Due to the optical theorem when the forward ${\\bf q}=0$ limit is taken this piece in \nthe kernel corresponds to the contribution to multiparticle production from on-shell gluons in the $s$-channel.} \nhas the following structure\n\\begin{equation}\nK_R (\\boldsymbol {p}_1,\\, \\boldsymbol {p}_2;\\, \\boldsymbol {p}_3,\\, \\boldsymbol {p}_4) = -N_cg^2\\left[ \\left( \\boldsymbol {p_3}+\\boldsymbol {p_4} \\right)^2 - \\frac{\\boldsymbol {p}_2^2\\boldsymbol {p}_4^2}{(\\boldsymbol {p}_2+\\boldsymbol {p}_3)^2}- \\frac{\\boldsymbol {p}_1^2\\boldsymbol {p}_3^2}{(\\boldsymbol {p}_1+\\boldsymbol {p}_4)^2} \\right]. \\label{eq:KernelR}\n\\end{equation}\nThis notation, with $p_1,\\, \\ldots ,\\, p_4$ being the cyclically ordered reggeized gluon momenta taken as incoming, \nwill be convenient for the generalization of this vertex to the $2 \\rightarrow m$ reggeized gluon transition case \nas we will see below. \n\nThe gluon Regge trajectory reads\n\\begin{equation}\n\\omega (\\boldsymbol {q}^2) = -\\frac{g^2N_c}{16 \\pi ^3 }\\int d^2\\boldsymbol {k}' \\frac{\\boldsymbol {q}^2}{\\boldsymbol {k}'^2(\\boldsymbol {k}'-\\boldsymbol {q})^2} \\ . \n\\label{eq:trajectoryreggegluon}\n\\end{equation}\nThe trajectory is IR divergent, requiring it to be regularized in general.\n\nWe will now show that the BFKL equation in Eq.~\\eqref{eq:BFKLconK} exhibits formally a dual $SL(2,C)$ symmetry, which, \nin contrast with the original $SL(2,C)$ symmetry of BFKL, uncovered by Fourier transforming into a coordinate \nrepresentation, is realized in the transverse momentum space. This new symmetry is closely analogous to the dual \nconformal symmetry observed in $\\mathcal{N}=4$ SYM for gluon scattering amplitudes, and we will see that it turns out \nto be broken by infrared effects just as in the four dimensional gauge theory.\n\nLet us now rewrite Eq.~\\eqref{eq:BFKLconK} in terms of dual variables. Taken as incoming, the external momenta are \n$\\boldsymbol {k}_A$, $-\\boldsymbol {k}_A+ \\boldsymbol {q}$, $\\boldsymbol {k}_B - \\boldsymbol {q}$ and $-\\boldsymbol {k}_B$ so, introducing the notation \n$x_{i,j}\\equiv x_i - x_j$, we define the new set of variables as\n\\begin{equation}\n\\boldsymbol {p}_1 = x_{1,2} = \\boldsymbol {k}_A, \\, \\boldsymbol {p}_2 = x_{2,3} = \\boldsymbol {q} -\\boldsymbol {k}_A, \\,\n\\boldsymbol {p}_3 = x_{3,4} = \\boldsymbol {k}_B - \\boldsymbol {q}, \\, \\boldsymbol {p}_4=x_{4,1} = -\\boldsymbol {k}_B. \n\\end{equation}\nEquivalently, we could have written $\\boldsymbol {k}_A = x_{1,2}, \\, \\boldsymbol {k}_B = x_{1,4}, \\, \n\\boldsymbol {q} = x_{1,3}$ with $x_1$ then being a simple shift of the origin for the external momenta.\n\nIn these new variables the gluon Regge trajectory is\n\\begin{equation}\n\\omega (\\boldsymbol {k}_A^2) = \\omega (x^2_{1,2})=-\\frac{g^2N_c}{16 \\pi ^3 }\\int d^2 x_I \\frac{x^2_{1,2}}{x^2_{I,1}x^2_{I,2}} \\ , \\label{eq:trajectoryreggegluonk1x}\n\\end{equation}\nwhere we have introduced $x_I$ through $\\boldsymbol {k}' = x_{I,2}$. Ignoring for the moment that this expression is divergent and thus ill-defined, we see that is has a formal two-dimensional conformal symmetry. It is formally invariant under translations, rotations and scalings of the $x_i$, and also under the conformal inversions $x_i \\rightarrow \\frac{x_i}{x_i^2}$, since they would imply\n\\begin{equation}\nd^2 x_I \\rightarrow \\frac{d^2 x_I}{x_I^4} \\ , \\;\\; x_{i,j}^2 \\rightarrow \\frac{x_{i,j}^2}{x_i^2x_j^2} \\ . \\label{eq:transfxij} \n\\end{equation}\nIn the same way $\\omega ((\\boldsymbol {k}_B-\\boldsymbol {q})^2) = \\omega (x^2_{2,3})$ is also formally conformally invariant. \nNow, given that the trajectory is infrared divergent one would expect this symmetry to be broken by the introduction \nof a regulator, an issue which is discussed in the next section.\n\nRewriting the full kernel~\\eqref{eq:Kernel} in terms of the $x_i$, with $\\boldsymbol {k}'=x_{1,I}$, we get\n\\begin{equation}\nK(x_{1,2},\\, x_{3,2};\\, x_{1,I},\\, x_{3,I}) = \n\\frac{K_R (x_{1,2},\\, x_{2,3}; x_{3,I}, \\, x_{I,1})}{8\\pi^3x_{1,2}^2x_{I,3}^2} \n+ \\left[ \\omega (x_{1,2}^2) + \\omega (x_{2,3}^2) \\right]\\delta ^{(2)}(x_{2,I}) \\ , \\label{eq:Kernelx}\n\\end{equation}\nwhere\n\\begin{equation}\nK_R (x_{1,2},\\, x_{2,3}; x_{3,I}, \\, x_{I,1}) = -N_cg^2\\left[ x_{1,3}^2 - \\frac{x_{2,3}^2 x_{I,1}^2}{x_{2,I}^2}- \\frac{x_{1,2}^2x_{I,3}^2}{x_{2,I}^2} \\right] \\ . \\label{eq:KernelRx}\n\\end{equation}\nUsing that $\\delta ^{(2)}(x_{2,I}) \\rightarrow x_2^2x_I^2 \\delta ^{(2)}(x_{2,I})$ under conformal inversions one then finds immediately that the kernel transforms covariantly \\footnote{This is again similar to the (conjectured) dual conformal symmetry of scattering amplitudes in ${\\cal N}=4$ SYM, under which the amplitudes transform covariantly, as opposed to the ordinary conformal symmetry which leave them invariant.}\n\\begin{equation}\nK(x_{1,2},\\, x_{3,2};\\, x_{1,I},\\, x_{3,I}) \\rightarrow x^2_2x^2_I K(x_{1,2},\\, x_{3,2};\\, x_{1,I},\\, x_{3,I}) \\ .\n\\end{equation}\nTogether with translations, rotations and dilatations this forms a dual $SL(2,C)$ symmetry, different from the one \npreviously known. More precisely, dilatations and rotations coincide with the original $SL(2,C)$-symmetry, while \ntranslations and inversions will be different. \n\nApplied to the BFKL equation~\\eqref{eq:BFKLconK} and using that the integration measure transforms according \nto~\\eqref{eq:transfxij} one finds that a factor of $\\frac{x_2^2}{x_I^2}$ is produced inside the integral. \nConsequently, if the Green function $F(\\omega , \\, x_{1,2},\\, x_{1,4},\\, x_{1,3})$ were to produce a factor of \n$x_2^2$ upon inversion, then its convolution with the kernel $K \\otimes F$, would transform in the same way as $F$ \nitself. Now, at lowest order, $F$ is simply given by the delta function, which indeed transforms in this way, \n$\\delta^{(2)} (\\boldsymbol {k}_A - \\boldsymbol {k}_B)=\\delta ^{(2)} (x_{2,4}) \\rightarrow x_2^2 x_4^2 \\delta ^{(2)} (x_{2,4})$. \nSince $F$ can be constructed through iterated convolution with the kernel, it follows that the Green function \nshould have the same conformal properties as the delta function. \n\nWe can obtain a formal expression for the Green function having the correct conformal properties by iteration. \nIntroducing the short-hand notation\n\\begin{equation}\n\\omega_0 \\left(\\bf{k}_A,\\, \\bf{q} \\right) \\equiv \\omega (\\boldsymbol {k}_A^2) + \\omega ((\\boldsymbol {k}_A-\\boldsymbol {q})^2), \n\\xi \\left(\\bf{k},\\, \\bf{k}_A,\\, \\bf{q}\\right) \\equiv \n\\frac{K_R (\\boldsymbol {k}_A,\\, \\boldsymbol {k}_A-\\boldsymbol {q}; -\\boldsymbol {k}+\\boldsymbol {q}, \\, -\\boldsymbol {k})}{8\\pi^3 \\boldsymbol {k}_A^2(\\boldsymbol {k} - \\boldsymbol {q})^2} \\ \n\\end{equation}\none finds (with $\\bf{k}_0 \\equiv \\bf{k}_A$):\n\\begin{eqnarray}\nF \\left(\\omega,\\bf{k}_A,\\bf{k}_B,\\bf{q}\\right) =\n{\\delta^{(2)} \\left(\\bf{k}_A-\\bf{k}_B\\right) + \\sum_{n=1}^\\infty \\prod_{i=1}^n \n\\int d^2 \\bf{k}_i \\, {\\xi \\left(\\bf{k}_i,\\bf{k}_{i-1},\\bf{q}\\right) \\over \n\\omega - \\omega_0 \\left(\\bf{k}_i,\\bf{q}\\right)} \\delta^{(2)}\n\\left(\\bf{k}_n-\\bf{k}_B\\right) \\over \\omega - \\omega_0 \\left(\\bf{k}_A,\\bf{q}\\right)}. \n\\end{eqnarray}\n\nRather than $\\omega$ it is more natural to use the rapidity difference, $Y$, between the external \nparticles as the evolution variable. To this end we perform the inverse Mellin transform\n\\begin{eqnarray}\n{\\cal F} \\left(\\bf{k}_A,\\bf{k}_B,\\bf{q},Y\\right) &=&\n\\int_{a-i \\infty}^{a+i \\infty} {d \\omega \\over 2 \\pi i} e^{\\omega Y}\nF \\left(\\omega,\\bf{k}_A,\\bf{k}_B,\\bf{q}\\right).\n\\end{eqnarray}\nThe formula $\\int_{a-i \\infty}^{a+i \\infty} {d \\omega \\over 2 \\pi i} e^{\\omega Y}\n\\prod_{i=0}^n \\frac{1}{\\omega-\\omega_i} = e^{\\omega_0 Y} \\prod_{i=1}^n\n\\int_0^{y_{i-1}} d y_i e^{\\omega_{i,i-1} y_i}$ for $n > 0$, with $\\omega_{i,j} \\equiv \n\\omega_i - \\omega_j, y_0 \\equiv Y$, is useful to obtain the final expression, written in dual $x$-variables:\n\\begin{equation}\n{\\cal F} \\left(x_{12}, x_{14},x_{13},Y\\right) = e^{\\omega _{2,1} Y} \\Bigg\\{\\delta^{(2)} \\left(x_{24}\\right)\n+ \\sum_{n=1}^\\infty \\prod_{i=1}^n \\int_0^{y_{i-1}} \\hspace{-0.3cm} d y_i\n\\int d^2 x_i \\, \\xi_{i,i-1} e^{\\omega_{i,i-1} y_i} \\delta^{(2)}\n\\left(x_{4,n} \\right) \\Bigg\\}, \n\\label{eq:Frapidityx}\n\\end{equation}\nwhere\n\\begin{eqnarray}\n\\omega_{i,i-1} &=& \\omega _0 \\left(x_{1,i},x_{13}\\right)-\\omega_0 \\left(x_{1,i-1},x_{13}\\right) \\ ,\\\\\n\\xi_{i,i-1} &=& {{\\bar \\alpha}_s \\over 2 \\pi} \\Bigg\\{{ x^2_{i,3} x^2_{i-1,1} +\nx^2_{i-1,3} x^2_{i,1} - x^2_{i-1,i} x^2_{13}\\over x^2_{i,3} x^2_{i-1,i} x^2_{i-1,1}}\\Bigg\\} \\ .\n\\end{eqnarray}\nThis representation preserves the transformation properties of the original equation. In the forward case, where the \nmomentum transfer is zero, the same structure remains with \n\\begin{equation}\n\\omega_{i,i-1} = 2\\left(\\omega \\left(x_{1,i}\\right)-\\omega \\left(x_{1,i-1}\\right)\\right) \\ ,\n\\xi_{i,i-1} = {{\\bar \\alpha}_s \\over \\pi} { 1 \\over x^2_{i-1,i}} \\ .\n\\end{equation}\nIn this case the solution also has a formal dual $SL(2,C)$ covariance. This should be contrasted with \nthe original $SL(2,C)$-invariance of the BFKL kernel, which does not appear in the forward case.\n\nBefore ending this section, it is noteworthy to mention that this formal dual $SL(2,C)$ covariance is \npresent in the same form for all color projections in the $t$-channel since they only differ by a different factor in front of $K_R$: with $N_c=3$, $c_1 = 1,\\, c_{8_a} = c_{8_s} = 1\/2,\\, c_{10} = c_{\\overline{10}} = 0$, etc.\n\n\\section{The effect of IR divergences}\n \nIn ${\\cal N}=4$ SYM infrared divergences break the dual conformal symmetry. For BFKL, such divergences cancel, opening the \npossibility that the dual $SL(2,C)$-symmetry remains exact. However, this turns out not to be the case. Perhaps the \nsimplest way to see this is by studying the forward case. If $F$ has the transformation properties of the delta \nfunction it can be written as\n\\begin{equation}\nF = F_1 \\delta ^{(2)}(\\boldsymbol {k}_A - \\boldsymbol {k}_B) + \\frac{1}{(\\boldsymbol {k}_A - \\boldsymbol {k}_B)^2}F_2 \\ ,\n\\end{equation}\nwhere $F_1$ and $F_2$ are dual conformally invariant, since $(\\boldsymbol {k}_A - \\boldsymbol {k}_B)^{-2}$ is the only other function \nthat transforms correctly. When $\\boldsymbol {q}=0$, $x_1=x_3$ and no non-trivial conformal invariant can be formed from the \nthree remaining $x_i$. $F_2$ can thus only be a function of $\\omega$ (or equivalently the rapidity $Y$), and the \ncoupling. But when forming physical quantities one integrates over $\\boldsymbol {k}_A$ and $\\boldsymbol {k}_B$ and the divergences \nat $\\boldsymbol {k}_A=\\boldsymbol {k}_B$ must cancel between $F_1$ and $F_2$. The factor $(\\boldsymbol {k}_A - \\boldsymbol {k}_B)^{-2}$ is singular \nenough to cancel one factor of the trajectory, but $F_1$ is obtained by repeated application of the trajectory \npart of the kernel so, starting from the second iteration, products of two or more trajectories will appear and \nthe divergences will fail to cancel.\n\nOne can also observe the breakdown of the dual $SL(2,C)$ symmetry directly by regularizing the integrals and \ncancelling the divergences explicitly when performing the iteration. One then finds that the first iteration \nrespects the symmetry, while the second iteration produces a contribution to $F_2$ (when $\\boldsymbol {q}=0$) proportional \nto an anomalous factor of the form $\\ln \\left( \\frac{(\\boldsymbol {k}_A-\\boldsymbol {k}_B)^4}{\\boldsymbol {k}_A^2 \\boldsymbol {k}_B^2}\\right)$, which \nbreaks the symmetry under inversions. The origin of this factor is the regularization of infrared divergences. \nFor example, using dimensional regularization with $D=4-2\\epsilon$\n\\begin{equation}\n\\omega (x^2_{1,2})=-\\frac{g^2N_c}{16 \\pi ^3 }(4 \\pi \\mu)^{2\\epsilon}\\int d^{2-2\\epsilon} x_I \\frac{x^2_{1,2}}{x^2_{I,1}x^2_{I,2}}\\approx -\\frac{g^2N_c}{8 \\pi ^2 }(4\\pi e^{-\\gamma})^\\epsilon \\left( \\ln \\frac{x^2_{1,2}}{\\mu ^2}- \\frac{1}{\\epsilon} \\right) \\ .\n\\end{equation}\nThe divergences will cancel between the trajectories and the real emission part of the kernel, but factors such \nas $\\ln x^2_{1,2}$ will add up giving a non-vanishing anomalous term. So, even though BFKL is infra-red finite, \na remnant of the divergences remains in the form of the breaking of the dual $SL(2,C)$ symmetry in the form given here.\n\nFurther insight can be gained by studying a standard representation of the Green function in the forward case, \nobtained by diagonalizing the BFKL kernel. It is\n\\begin{equation}\n{\\cal F} \\left(x_{1,2},x_{1,4},Y\\right) = \n\\sum_{n=-\\infty}^\\infty \\int \\frac{d \\gamma}{2 \\pi i}\n\\left({x^2_{1,2} \\over x^2_{1,4}}\\right)^{\\gamma-\\frac{1}{2}}\n\\frac{e^{{\\bar{\\alpha}_s}\\chi_n\\left(\\gamma\\right) Y + i n \\theta_{2,4}}}{\\pi\\sqrt{x_{1,2}^2 x_{1,4}^2}}\\ ,\n\\end{equation}\nwith $\\chi_n \\left(\\gamma\\right) = 2 \\Psi\\left(1\\right)- \\Psi\\left(\\gamma+\\frac{|n|}{2}\\right)\n- \\Psi\\left(1-\\gamma+\\frac{|n|}{2}\\right)$ and \n$\\cos{\\theta_{2,4}} = { x_{1,2} \\cdot x_{1,4} \\over \\sqrt{x^2_{1,2} x^2_{1,4}}}$.\nIn this representation any dependence on an IR cutoff has canceled explicitly, and one can check that the covariance under conformal inversions is lost.\n\nAn important issue is whether the dual $SL(2,C)$ symmetry is broken beyond repair or whether it can be deformed to take into consideration the anomalous terms. In a best case scenario the symmetry would obey to all orders a simple relation such as the anomalous Ward identity satisfied by the dual conformal symmetry of $\\mathcal{N}=4$ scattering amplitudes \\cite{confward}. This issue is studied in \\cite{johan}, with the result that the dual $SL(2,C)$ does not obey such a simple all-order relation, but can still be deformed so that it becomes exact, at least up to the order studied. The representation then becomes coupling-dependent, but encouragingly, it seems to do so in such a way that the algebra generated by the original and dual $SL(2,C)$ symmetries remains coupling-independent.\n\n\\section{$2\\rightarrow m$ reggeized gluon vertex}\n\n\nThe BFKL amplitude will violate bounds imposed by unitarity at sufficiently high energies. \nIn order to restore unitarity, one of the new elements that must be introduced is a vertex in which the number \nof reggeized gluons in the $t$-channel is not conserved. As shown in Fig.~\\ref{fig:vertex} we choose to write this \n$2\\rightarrow m$ vertex (see, for example, Eq.~(3.57) of \\cite{BartelsEwerz}) using a convenient assignment \nof the momentum indices\n\\begin{align}\n&K_{2\\rightarrow m}^{\\{ b \\} \\rightarrow \\{a \\}}(\\boldsymbol {p}_2,\\, \\boldsymbol {p}_3;\\, \\boldsymbol {p}_4,\\, \\ldots ,\\, \\boldsymbol {p}_{m+2},\\, \\boldsymbol {p}_1 ) = f_{a_1b_1c_1}f_{c_1a_2c_2}\\cdots f_{c_{m-1}a_mb_2} g^m \\nonumber \\\\\n& \\times \n\\left[ (\\boldsymbol {p}_4 + \\cdots + \\boldsymbol {p}_1)^2 - \\frac{\\boldsymbol {p}_3^2(\\boldsymbol {p}_5 + \\cdots + \\boldsymbol {p}_1)^2}{(\\boldsymbol {p}_3 + \\boldsymbol {p}_4)^2}\n - \\frac{\\boldsymbol {p}_2^2(\\boldsymbol {p}_4 + \\cdots + \\boldsymbol {p}_{m+2})^2}{(\\boldsymbol {p}_1 + \\boldsymbol {p}_2)^2} + \\frac{\\boldsymbol {p}_1^2\\boldsymbol {p}_3^2(\\boldsymbol {p}_5 + \\cdots + \\boldsymbol {p}_{m+2})^2}{(\\boldsymbol {p}_1 + \\boldsymbol {p}_2)^2(\\boldsymbol {p}_3 + \\boldsymbol {p}_4)^2} \\right] \\label{eq:K2m} \\ ,\n\\end{align}\nwhere the $a_1,\\, b_1$ etc. are the color indices of the reggeized gluons and $f_{ijk}$ the structure constants of \n$SU(N_c)$. \n\n\\begin{figure}[ht]\n\\psfrag{x1}{$x_1$} \\psfrag{x2}{$x_2$} \\psfrag{x3}{$x_3$} \\psfrag{x4}{$x_4$} \\psfrag{x5}{$x_5$} \\psfrag{p1}{$\\boldsymbol {p}_1$} \\psfrag{p2}{$\\boldsymbol {p}_2$} \\psfrag{p3}{$\\boldsymbol {p}_3$} \\psfrag{p4}{$\\boldsymbol {p}_4$} \\psfrag{p5}{$\\boldsymbol {p}_5$} \\psfrag{pmp2}{$\\boldsymbol {p}_{m+2}$} \\psfrag{a1}{$a_1$} \\psfrag{a2}{$a_2$} \\psfrag{amm1}{$a_{m-1}$} \\psfrag{am}{$a_m$} \\psfrag{b1}{$b_1$} \\psfrag{b2}{$b_2$}\n\\begin{center}\n\\includegraphics{reggluonvertex.eps}\n\\caption{\\small The $2\\rightarrow m$ reggeized gluon vertex. All momenta are taken as ingoing.} \\label{fig:vertex}\n\\end{center}\n\\end{figure}\n\nWritten in terms of $x$ variables this becomes\n\\begin{align}\n&K_{2\\rightarrow m}^{\\{ b \\} \\rightarrow \\{a \\}}(x_{23},\\, x_{34};\\, x_{45},\\, \\ldots ,\\, x_{m+2,1},\\, x_{12}) = \\nonumber \\\\ &f_{a_1b_1c_1}f_{c_1a_2c_2}\\cdots f_{c_{m-1}a_mb_2} g^m \\left[ x_{24}^2 - \\frac{x_{34}^2x_{25}^2}{x_{35}^2} -\\frac{x_{23}^2x_{14}^2}{x_{13}^2} +\\frac{x_{23}^2x_{34}^2x_{15}^2}{x_{13}^2 x_{35}^2} \\right] \\ ,\n\\end{align}\nand is manifestly conformally covariant. The assignment of the momenta in \\eqref{eq:K2m} was chosen so that \nthe vertex takes a form independent of $m$ when written in terms of the $x_i$. Note that the last term vanishes \nwhen $m=2$ since then $x_1 = x_5$, and one recovers the corresponding term in the BFKL kernel. \n\nIn \\cite{reggevertex2a4} it was shown that the $2\\rightarrow 4$ reggeized gluon vertex exhibited the same coordinate representation $SL(2,C)$-invariance as the BFKL equation. This was taken to indicate that a unitary, two-dimensional CFT describing scattering amplitudes in the Regge limit should exhibit this $SL(2,C)$-invariance. Our results would seem to indicate that such a theory should also be covariant under the dual $SL(2,C)$.\n\n\\section{Conclusions}\n\n\nWe have shown that not only does the LLA BFKL kernel, and its extension in the form of the $2 \\rightarrow m$ reggeized gluon vertex, exhibit the ordinary $SL(2,C)$-symmetry, found by Lipatov but also a dual $SL(2,C)$, analogous to the dual conformal symmetry of $\\mathcal{N}=4$. It is tempting to interpret these symmetries as reductions to the transverse plane of the conformal and dual conformal symmetries of the supersymmetric theory, although it is not clear exactly how such a reduction should be carried out. Purely transverse versions of the conformal algebras are not symmetries of the 4-dimensional gauge theory amplitudes, but seem to emerge in the Regge limit. \n\nAlso, the dual invariance of the reggeized gluon vertex suggests that a unitary two-dimensional CFT describing high-energy gauge theory should have both $SL(2,C)$ groups. In future work, having identified the dual $SL(2,C)$ one can try to understand the origin of the integrability of the Regge limit in terms of the integrability of $N=4$ SYM.\n\n\n\n\\vspace{5mm}\n\\centerline{\\bf Acknowledgments}\n\nWe would like to thank Lev Lipatov for useful discussions. The work of C. G. has been partially\nsupported by the Spanish DGI contract FPA2003-02877 and the CAM grant HEPHACOS\nP-ESP-00346. The work of J. G. is supported by a Spanish FPU grant.\n\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\nIn this paper, we are concerned with the $L^p$ mapping properties of \nthe pseudodifferential operators in the form \n\\begin{equation}\n\\label{eq:1}\nT_\\sigma f(x)=\\int\\limits_{{\\mathbf R}^n} \\sigma(x,\\xi) e^{2\\pi i \\xi x} \\hat{f}(\\xi) d\\xi.\n\\end{equation} \nThe operators $T_\\sigma$ have been subject of continuous interest \nsince the sixties. We should mention that their usefullness in \n the study of partial differential equations have been realized much \nearlier, but it seems that their systematic study began with the \nfundamental works of Kohn and Nirenberg, \\cite{KN} and H\\\"ormander, \n\\cite{Hor}. \n\nTo describe the \nresults obtained in these early papers, \ndefine the H\\\"ormander's class $S^m$, which consists \nof all functions $\\sigma(x,\\xi)$, so that \n\\begin{equation}\n\\label{eq:2}\n|D^\\beta_x D^\\alpha_\\xi \\sigma(x, \\xi)|\\leq C_{\\alpha, \\beta} (1+|\\xi|)^{m-|\\alpha|}.\n\\end{equation}\nfor all multiindices $\\alpha, \\beta$. \nA classical theorem in \\cite{Hor} then states that \n$Op(\\sigma):H^{s+m,p}\\to H^{s, p}$ for all $s\\geq 0$ and \n$1
^{|\\alpha|} \n|D_\\xi^\\alpha [\\sigma(x+y, \\xi)-\\sigma(x, \\xi)]| \\leq C_\\alpha \\omega(|y|)\n$$\nand assume that $\\sum_{j>0} \\omega(2^{-j})^2<\\infty$. Then for \nall $1
0} \nC^\\gamma S^0\\subset C^\\omega S^{0}_{1,0}$. \nRelated results can be found in the work of M. Taylor, \\cite{Taylor} (see Proposition 2.4, p. 23) and J. Marschall, \\cite{Marschall} where the \nspaces $C^\\omega$ are replaced by $H^{\\varepsilon,p}$ spaces with $p$ as large as one wish and $0<\\varepsilon=\\varepsilon(p)<<1$ (see also \\cite{Taylor}, p. 61)\n\nOne of the purposes of this work is to get away from the \ncontinuity requirements on $x\\to \\sigma(x, \\xi)$. Even more importantly, \nwe would like to replace the pointwise conditions on the derivatives of $\\xi$ \nby averaged ones. This particular point has not been thoroughly \nexplored appropriately \n in the literature in the author's opinion, \nsee Theorem \\ref{theo:1} below. \n\nOn the other hand, a particular motivation for such considerations is \nprovided by the recent papers of Rodnianski-Tao \\cite{RT} and \nthe author \\cite{Stefanov}, where concrete parametrices \n(i.e. pseudodifferential operators, representing \napproximate solutions to certain PDE's) \nwere constructed for the \nsolutions of certain first order perturbation of the wave \nand Schr\\\"odinger equations. \nA very quick inspection of these examples shows that\\footnote{Most \nreaders are \n likely to have their own fairly long list \nwith favorite examples, for which \n the H\\\"ormander condition fails.} \n {\\it \nthey do not obey\npointwise conditions on the derivatives on the \nsymbols} and thus, these methods fail \nto imply $L^2$ bounds for \nthese (and related problems). Moreover, one often times has to deal with the \nsituation, where the maps $\\xi\\to \\sigma(x, \\xi)$ \nare not smooth in a pointwise sense. \nOn the other hand, one may still be able to control averaged quantities like \n\\begin{equation}\n\\label{eq:5}\n\\sup_x \\norm{\\sigma(x, \\xi)}{H^{n\/2}_\\xi}<\\infty.\n\\end{equation}\nThis will be our treshold condition for $L^2$ boundedness, \nwhich we try to achieve. \\\\ Heuristically at least, \\eqref{eq:5} \nmust be ``enough'' in some sense, \n since if we had simple symbols like $\\sigma(x,\\xi)=\\sigma_1(x) \\sigma_2(\\xi)$, \nthen the $L^2$ boundedness of $Op(\\sigma)$ is equivalent to \n $\\norm{\\sigma_1}{L^\\infty_x}<\\infty, \n\\norm{\\sigma_2}{L^\\infty_\\xi}<\\infty$. Clearly, \n$\\norm{\\sigma_2}{L^\\infty_\\xi({\\mathbf R}^n)}$ just fails to be controlled by \n\\eqref{eq:5}, but on the other hand, the quanitity in \\eqref{eq:5} \nis controlled by the \nappropriate Besov space $B^{n\/2}_{2,1}$ norm. \n\n A final motivation for \nthe current study is to achieve a scale\n invariant condition, which gives an estimate of the \n$L^2\\to L^2$ ($L^p\\to L^p$) norm of \n$Op(\\sigma)$ in terms of a {\\it scale invariant quantity}, \nthat is, we aim at showing \nan estimate, \n$$\n\\norm{Op(\\sigma)}{L^p\\to L^p}\\leq C \\norm{\\sigma}{Y} \\norm{f}{L^p}, \n$$\nwhere for every $\\lambda\\neq 0$, one has $\\norm{\\sigma(\\lambda \\cdot,\\lambda^{-1} \n\\cdot) }{Y}=\\norm{\\sigma}{Y}$. \n\nIn that regard, note that the condition (which is one of the requirements of \n the H\\\"ormander class $S^0$) \n\\begin{equation}\n\\label{eq:6}\n\\sup_{x}| D_\\xi^\\alpha \\sigma(x, \\xi)|\\leq C_\\alpha |\\xi|^{-|\\alpha|}\n\\end{equation}\nis scale invariant in the sense described above. Moreover, \nby the standard Calder\\'on-Zygmund theory (see \\cite{Stein}), \nthe pointwise condition \\eqref{eq:6} together with \n$\\|T_\\sigma\\|_{L^2\\to L^2}<\\infty$ implies \n$$\nT_\\sigma f(x)=\\int K(x, x-y) f(y) dy,\n$$\nwhere $K(x,\\cdot)$ satisfies the H\\\"ormander-Mihlin conditions, namely $|K(x, z)|\\leq C|z|^{-n}$ and \n$|\\nabla_z K(x,z)|\\leq C|z|^{-n-1}$, where the constant $C$ depends on \nthe constants \n$C_\\alpha: |\\alpha|<[n\/2]+1$ in \\eqref{eq:6}. \nThis in turn is enough to conclude \nthat $T_\\sigma:L^p \\to L^p$ for all $1
2$, \nthere exists $\\sigma(x,\\xi)$ so that $\\sup_{x} | D_\\xi^{\\alpha}\\sigma(x, \\xi)|\\leq C_\\alpha |\\xi|^{-|\\alpha|}$ and $\\sup_x \\norm{\\sigma(x, \\cdot)}{W^{p, n\/p}}<\\infty$, \nbut $T_\\sigma$ fails to be bounded on $L^2({\\mathbf R}^n)$. \n\\end{theorem}\n{\\bf Remark:} \n\\begin{enumerate} \n\\item Note that the estimate on $T_\\sigma$ is scale invariant. \n\\item The sharpness claim of the theorem, roughly speaking, \nshows that in the scale of spaces\\footnote{Note that \nthese spaces scale the same and moreover \nby Sobolev embedding these are strictly decreasing sequence, at least \nfor $2\\leq p<\\infty$.} $W^{p, n\/p}$, $\\infty\\geq p\\geq 2$, one may \nnot require anything less than $W^{2,n\/2}=H^{n\/2}$ of the symbol in \norder to ensure $L^2$ boundedness.\n\\item The counterexample to which we refer in Theorem \\ref{theo:1} \nis a simple variation of the well-known example of $\\sigma\\in \nS^{0}_{1,1}$, the ``forbidden class'', \n which fails to be $L^2$ bounded, see \\cite{Stein}, p. 272 and Section \n\\ref{sec:counter} below. \n \\end{enumerate}\nOur next result concerns $L^p$ boundedness for $T_\\sigma$. \n\\begin{theorem}($L^p$ bounds) \n\\label{theo:3}\nFor the pseudodifferential operator $T_\\sigma$ there \nis the estimate for all $2
1$, there exists a homogeneous of degree zero symbol \n$\\sigma(x, \\xi):\\mathbf R^2\\times \\mathbf R^2\\to \\mathbf R^1$, \nso that $\\sup_{x, \\xi} |\\sigma(x, \\xi)|<\\infty$ and \n$\\sup_x \\norm{\\sigma(x,\\xi)}{W^{1, 1}(\\mathbf S^1)}<\\infty$, \nand so that $\\norm{T_\\sigma}{L^2\\to L^2}>N$. \n\\end{proposition}\nThe counterexample considered here is a smoothed out version of the \nmaximal directional Hilbert transform in the plane $H_* f(x)=\\sup_{u\\in \\mathbf S^1} \n|H_u f(x)|$. We mention the spectacular recent result of \nLacey and Li, \\cite{LL} showing the boundedness of \n$H_*$ on $L^p(\\mathbf R^2): 2
0$, define \n$$\n\\mathcal C_\\delta f(x)=\\sup_{u>0} \\int_{\\mathbf R^1} (1-\\xi^2\/u^2)^{\\delta}_+ \ne^{2\\pi i \\xi x} \\hat{f}(\\xi) d\\xi.\n$$ \nClearly, as a limit as $\\delta\\to 0$, we get the Carleson's operator. \nUnfortunately, one cannot conclude that \n$\\sup_{\\delta>0}\\|C_\\delta\\|_{L^p}<\\infty$, for\nthat would imply the famous Carleson-Hunt theorem. \nOn the other hand, define the maximal ''thin\ninterval operator''\n$$\nT_m f(x)=\\sup_{u>0} \\int_{\\mathbf R^1} \\varphi(2^m(1-\\xi^2\/u^2)) \ne^{2\\pi i \\xi x} \\hat{f}(\\xi) d\\xi.\n$$\nA \nsimple argument based on (the proof of) Theorem \\ref{theo:3} yields \n\\begin{proposition}\n\\label{prop:Carl}\nFor any $\\varepsilon>0, 1
0} \\int_{{\\mathbf R}^n} (1-|\\xi|^2\/u^2)^{\\delta}_+ \ne^{2\\pi i \\xi x} \\hat{f}(\\xi) d\\xi.\n$$\n{\\it in any dimension}. \n\\end{itemize}\n\\begin{proof}\nIt clearly suffices to show the pointiwise estimate \n$|T_m f(x)|\\leq C M (\\sup_k P_k f)(x)$ for any $k$.\nThe statements about $B^0_{p,1}\\to L^p$ bounds follow by elementary Littlewood-Paley theory and\nthe $l^p$ bounds for the Hardy-Littlewood maximal function. The restricted-to-weak estimate \n$F^0_{1,\\infty}\\to L^{1,\\infty}$ for $C_\\delta$ follows by summing an \n{\\it exponentially decaying} series in\nthe quasi-Banach space $L^{1, \\infty}$. \n\nBy support considerations, it is clear that \n\\begin{eqnarray*}\nT_m f(x) & = &\\sum\\limits_k \\sup_{u>0} \\int_{\\mathbf R^1} \\varphi(2^m(1-\\xi^2\/u^2)) \ne^{2\\pi i \\xi x}\\varphi(2^{-k}\\xi) \\hat{f}(\\xi) d\\xi= \\\\\n&=&\n\\sum\\limits_k \\sup_{u\\in (2^{k-2}, 2^{k+2})} \\int_{\\mathbf R^1} \\varphi(2^m(1-\\xi^2\/u^2)) \ne^{2\\pi i \\xi x}\\varphi(2^{-k}\\xi) \\hat{f}(\\xi) d\\xi= \\\\\n&=& \\sum\\limits_k T_{m, u(\\cdot)\\in (2^{k-2}, 2^{k+2})} f_k.\n\\end{eqnarray*}\nClearly, the requirement $u\\in (2^{k-2}, 2^{k+2})$ creates (almost) \ndisjointness in the $x$ support, whence \n\\begin{equation}\n\\label{eq:819}\n|T_m f(x)|\\leq C \\sup\\limits_k |T_{m, u(\\cdot)\\in (2^{k-2}, 2^{k+2})} f_k(x)|.\n\\end{equation}\nOur basic claim is that \n\\begin{equation}\n\\label{eq:820}\n|T_{m, u(\\cdot)\\in (2^{k-2}, 2^{k+2})} f_k(x)|\\leq C M(f_k).\n\\end{equation}\nClearly \\eqref{eq:819} and \\eqref{eq:820} imply $\\sup_m |T_m f(x)|\\leq C M(\\sup\\limits_k |f_k|)$, \nwhence the Proposition \\ref{prop:Carl}. \\\\\nBy scale invariance, \\eqref{eq:820} reduces to the case $k=0$, that is we need to show \n$$\n|T_{m, u(x)\\in (1\/4, 4)} P_0 f(x)|\\leq C M (P_0 f)(x).\n$$\nfor any Schwartz function $f$ and any \n$m>>1$. By \\eqref{eq:32} (in the proof of Theorem \\ref{theo:3} below), it will suffice to\nshow \n\\begin{equation}\n\\label{eq:920}\n\\sum\\limits_l 2^l \\sup\\limits_x \\|P_l^\\xi [\\varphi(2^m(1-\\xi^2\/u(x)^2)) \\varphi(\\xi)]\\|_{L^1(\\mathbf R^1)}\\lesssim 1.\n\\end{equation}\nfor any measurable function $u$, which takes its values in $(1\/4, 4)$. \\\\\nFor \\eqref{eq:920}, we have \n\\begin{eqnarray*}\n& & \\sum\\limits_{l 0} \\f{D_\\xi^\\alpha q(\\theta_0)}{\\alpha!}(\\theta-\\theta_0)^\\alpha.\n\\end{equation}\nHere, $D_\\xi^\\alpha q(\\theta_0)$ should be understood as taking $\\alpha$ derivatives of \nthe corresponding homogeneous polynomial and evaluating at $\\theta_0$. \nThe following lemma is standard, but since we need a specific dependence of our \nestimates upon the parameter $\\alpha$, we state it here for completeness. \n\\begin{lemma}\n\\label{le:os}\nLet $q:\\mathbf S^{n-1}\\to \\mathcal C$ and $q_m=P^{\\xi\/|\\xi|}_m q$. Then, there is a constant $C_n$, so that \nfor every $1\\leq p\\leq \\infty$, there is a the estimate \n$$\n\\|D^\\alpha_\\xi q_m\\|_{L^p(\\mathbf S^{n-1})}\\leq C_n^{|\\alpha|} 2^{m |\\alpha|} \\norm{q_m}{L^p(\\mathbf S^{n-1})}.\n$$\n\\end{lemma}\nThe proof of Lemma \\ref{le:os} is standard. One way to proceed is to note that if we\nextend the function $q_m$ off $\\mathbf S^{n-1}$ to some annulus, say via \n$Q_m(\\xi)=\\varphi(|\\xi|) q_m(\\xi\/|\\xi|)$, and \nthen \n$$\n \\|D^\\alpha_\\xi q_m\\|_{L^p(\\mathbf S^{n-1})}\\lesssim \\|D^\\alpha_\\xi Q_m \\|_{L^p({\\mathbf R}^n)}\n$$\n\n\n\n\n\\section{$L^p$ estimates for PDO with rough symbols}\nWe start with the $L^2$ estimate to illustrate the main ideas in the proof. \n\\subsection{$L^2$ estimates: Proof of Theorem \\ref{theo:1}}\nOur first remark is that we will for convenience consider only \nreal-valued symbols $\\sigma$, since of course the general \ncase follows from splitting into a real and imaginary part. \n\nTo show $L^2$ estimates for $T_\\sigma$, it is equivalent to show $L^2$ \nestimates for the adjoint operator, which takes the \nform\\footnote{ There is the small technical problem \nthat the $\\xi$ integral does not converge \nabsolutely. \nThis can be resolved by judicious placement of cutoffs \n$\\chi(\\xi\/N)$, after which, one may subsume that \npart in $\\hat{f}(\\xi)$. In the end \nwe let $N\\to \\infty$ and all the estimates will be \nindependent of the cutoff constant $N$.} \n$$\nT_\\sigma^* g(x)=\\int e^{2\\pi i \\xi\\cdot x} (\\int e^{-2\\pi i \\xi\\cdot y} \n[g(y) \\sigma(y, \\xi)] dy )d\\xi.\n$$\nOur next task is to decompose $T_\\sigma^* g$ and we start by \ntaking a Littlewood-Paley partition of unity in the $\\xi$ variable for $g$. We have \n$$\nT_\\sigma^* g (x)=\\sum\\limits_{l\\in \\mathcal Z} \\int e^{2\\pi i \\xi\\cdot x} \n(\\int e^{-2\\pi i \\xi\\cdot y} \n[ g(y) P_l^\\xi \\sigma(y, \\xi)] dy )d\\xi\n$$ \nNow that the function $g$ is frequency localized at frequency $2^l$, \nwe introduce further decomposition in the \n$\\xi$ integration. \n\nFor the $L^2$ estimates, because of the orthogonality, \nwe only need rough partitions, so for each fixed $l$, \ntake a tiling of ${\\mathbf R}^n$ composed of cubes $\\{Q\\}$ \nwith diameter $2^{-l}$. Denote the characteristic functions of $Q$ by $\\chi_Q$. We have \n$$\nT_\\sigma^* g (x)=\\sum\\limits_{l\\in \\mathcal Z} \\sum\\limits_{Q:d(Q)=2^{-l}} \n\\int e^{2\\pi i \\xi\\cdot x} \\chi_Q(\\xi)\n(\\int e^{-2\\pi i \\xi\\cdot y} \n[ g(y) P_l^\\xi \\sigma(y, \\xi)] dy )d\\xi\n$$\nThe main point of our next decompositions is that the \nfunction $P_l^\\xi \\sigma$ is essentially constant in $\\xi$ \nover any fixed cube $Q$. We exploit that by observing that \n$\\xi\\to P_l^\\xi\\sigma(x, \\xi)$ is an entire function and there is \n the expansion \n$$\nP_l^\\xi \\sigma(y, \\xi) \\chi_Q(\\xi)=[P_l^\\xi \\sigma(y, \\xi_Q)+\\sum\\limits_{\\alpha: |\\alpha|>0}^\\infty\n\\f{D_\\xi^\\alpha P_l^\\xi \\sigma(y, \\xi_Q)}{\\alpha!} (\\xi-\\xi_Q)^\\alpha]\\chi_Q(\\xi)\n$$\nfor any fixed $y$ and for any $\\xi_Q\\in Q$. Note that \n$D_\\xi^\\alpha P_l^\\xi \\sigma(y, \\xi_Q)\\sim 2^{l|\\alpha|} P_l^\\xi \\sigma(y, \\xi_Q)$ and \n$ |(\\xi-\\xi_Q)^\\alpha|\\lesssim 2^{-l|\\alpha|}$, by support consideration \n(recall $d(Q)=2^{-l}$). On a heuristic level, by the presence of $\\alpha!$, \none should think that the series \nabove behave like $P_l^\\xi \\sigma(y, \\xi_Q)$ plus exponential tail. \n\nGoing back to $D_\\xi^\\alpha P_l^\\xi$, as we have mentioned in Section \\ref{sec:prelim}, we can write \n$D_\\xi^\\alpha P_l^\\xi=2^{l|\\alpha|} P_{l, \\alpha}^\\xi$, where \n$P_{l, \\alpha}^\\xi$ is given by the multiplier $\\varphi(2^{-l}\\xi) (2^{-l} \\xi)^\\alpha$. \nIt is clear that \n$\\|P_{l, \\alpha}^\\xi f\\|_{L^2({\\mathbf R}^n)}\\leq C_n^{|\\alpha|}\\|P_l^\\xi f\\|_{L^2({\\mathbf R}^n)}$. \n\nThus, we have arrived at \n\\begin{eqnarray*}\n& & T_\\sigma^* g (x)=\\sum\\limits_{|\\alpha|\\geq 0} \\sum\\limits_{l\\in \\mathcal Z} \\sum\\limits_{Q:d(Q)=2^{-l}} \n\\int e^{2\\pi i \\xi\\cdot x} \\chi_Q(\\xi) \n\\f{2^{l|\\alpha|}(\\xi-\\xi_Q)^\\alpha}{\\alpha!} \\times \\\\\n& & \\times \n(\\int e^{-2\\pi i \\xi\\cdot y} \n[ g(y) P_{l,\\alpha}^\\xi \\sigma(y, \\xi_Q)] dy )d\\xi= \n\\sum\\limits_{l,\\alpha}(\\alpha!)^{-1} \\sum\\limits_{l, Q:d(Q)=2^{-l}} P_{Q, l, \\alpha} [ g(\\cdot) \nP_{l,\\alpha}^\\xi \\sigma(\\cdot, \\xi_Q)], \n\\end{eqnarray*}\nwhere $P_{Q, l, \\alpha}$ acts via $\\widehat{P_{Q, l, \\alpha} f}(\\xi)= \\chi_Q(\\xi) \n2^{l|\\alpha|}(\\xi-\\xi_Q)^\\alpha \\hat{f}(\\xi)$. Note \n$$\n\\|P_{Q, l, \\alpha}\\|_{L^2\\to L^2}=\\sup_\\xi | \\chi_Q(\\xi) \n2^{l|\\alpha|}(\\xi-\\xi_Q)^\\alpha|\\leq 1. \n$$\n\nFor fixed $l, \\alpha$, take $L^2$ norm. Using the orthogonality of \n$P_{Q, l, \\alpha}$ and its boundedness on $L^2$, we obtain \n\\begin{eqnarray*}\n& & \\|\\sum\\limits_{Q:d(Q)=2^{-l}} P_{Q, l, \\alpha} [ g(\\cdot) \nP_{l,\\alpha}^\\xi \\sigma(\\cdot, \\xi_Q)]\\|_{L^2}^2= \n\\sum\\limits_{Q:d(Q)=2^{-l}} \\|P_{Q, l, \\alpha} [ g(\\cdot) \nP_{l,\\alpha}^\\xi \\sigma(\\cdot, \\xi_Q)]\\|_{L^2}^2\\leq \\\\\n& & \\leq \\sum\\limits_{Q:d(Q)=2^{-l}} \\|g(\\cdot) \nP_{l,\\alpha}^\\xi \\sigma(\\cdot, \\xi_Q)]\\|_{L^2}^2=\\int |g(y)|^2 \n(\\sum\\limits_{Q} |P_{l,\\alpha}^\\xi \\sigma(y, \\xi_Q)|^2) dy . \n\\end{eqnarray*}\nWe now again use $P_{l,\\alpha}^\\xi \\sigma(y, \\xi_Q)\\sim P_{l,\\alpha}^\\xi \\sigma(y, \\eta)$ for any\n$\\eta\\in Q$, this time to estimate the contribution of $\\sum_{Q} |P_{l,\\alpha}^\\xi \\sigma(y, \\xi_Q)|^2$. \nThis is done as follows. Expand \n\\begin{equation}\n\\label{eq:859}\nP_{l, \\alpha} ^\\xi \\sigma(y, \\xi_Q)= \\sum\\limits_{\\beta: |\\beta|\\geq 0}^\\infty\n\\f{D_\\eta^\\beta P_{l, \\alpha}^\\eta \\sigma(y, \\eta)}{\\beta!} (\\xi_Q-\\eta)^\\beta,\n\\end{equation}\nto be used for $\\eta\\in Q$. Thus, if we average over $Q$, \n\\begin{eqnarray*}\n& & |P_{l, \\alpha} ^\\xi \\sigma(y, \\xi_Q)|=\\left(|Q|^{-1} \\int_Q |\\sum\\limits_{\\beta: |\\beta|\\geq 0}^\\infty\n\\f{D_\\xi^\\beta P_{l, \\alpha}^\\xi \\sigma(y, \\eta)}{\\beta!} (\\xi_Q-\\eta)^\\beta|^2 d\\eta\n\\right)^{1\/2}\\leq \\\\\n& &\\leq |Q|^{-1\/2} \\sum\\limits_{\\beta: |\\beta|\\geq 0}^\\infty \\f{C_n^{|\\beta|} \n2^{-l|\\beta|}}{\\beta!} \\left(\\int\\limits_Q | D_\\xi^\\beta P_{l, \\alpha}^\\xi \\sigma(y, \\eta)|^2 d\\eta\\right)^{1\/2}.\n\\end{eqnarray*}\nand so (recalling $|Q|\\sim 2^{-l n}$)\n\\begin{eqnarray*}\n& &(\\sum\\limits_{Q} |P_{l,\\alpha}^\\xi \\sigma(y, \\xi_Q)|^2)^{1\/2}\\leq C 2^{l n\/2} \\sum\\limits_{\\beta} \\f{C_n^{|\\beta|} \n2^{-l|\\beta|}}{\\beta!} \\|D_\\xi^\\beta P_{l, \\alpha}^\\xi \\sigma(y, \\cdot)\\|_{L^2} \\leq \\\\\n& & \\leq \nC 2^{l n\/2} \\|P_{l, \\alpha}^\\xi \\sigma(y, \\cdot)\\|_{L^2}.\n\\end{eqnarray*}\nThus, \n\\begin{eqnarray*}\n& & \\|T_\\sigma^* g\\|_{L^2} \\lesssim \\sum\\limits_{l, \\alpha} 2^{l n\/2} (\\alpha!)^{-1} \\left(\\int |g(y)|^2 \n \\|P_{l, \\alpha}^\\xi \\sigma(y, \\cdot)\\|_{L^2}^2 dy\\right)^{1\/2}\\lesssim \\\\\n & &\\lesssim \n \\sum\\limits_{l, \\alpha} 2^{l n\/2} (\\alpha!)^{-1} \\|g\\|_{L^2} \\sup\\limits_y \\|P_{l, \\alpha}^\\xi \\sigma(y, \\cdot)\\|_{L^2}. \n\\end{eqnarray*}\nFurthermore, \n$$\n\\sup_y \\|P_{l,\\alpha}^\\xi \\sigma(y,\\cdot)\\|_{L^2}\\leq C_n^{|\\alpha|} \n\\sup_y \\|P_{l}^\\xi \\sigma(y,\\cdot)\\|_{L^2}.\n$$\nPut everything together \n\\begin{eqnarray*}\n& & \\|T_\\sigma^* g\\|_{L^2}\\leq C_n \n\\norm{g}{L^2}\\sum\\limits_{\\alpha} (\\alpha!)^{-1} C_n^{|\\alpha|} \n\\sum_l 2^{ln\/2} \n \\sup_y \\|P_{l}^\\xi \\sigma(y,\\cdot)\\|_{L^2} \n\\leq \\\\\n& & \\leq D_n \\norm{g}{L^2} \\sum_l 2^{ln\/2} \n \\sup_y \\|P_{l}^\\xi \\sigma(y,\\cdot)\\|_{L^2}, \n\\end{eqnarray*}\nas desired. \n\\subsection{$L^p$ estimates: $2 3$. Also since $P_k:L^1\\to L^1$, we get \n$$\n\\|\\widehat{\\psi^\\alpha_{l,Q}}\\|_{L^1}\\leq C_n^{|\\alpha|} \n\\|P_k[\\widehat{\\psi_{l,Q}}]\\|_{L^1}\\leq C_n^{|\\alpha|} \n\\|\\widehat{\\psi_{l,Q}}\\|_{L^1}\\leq \nC_n^{|\\alpha|}.\n$$\nThus, it remains to show for every $x$ and for {\\it any}\n$\\{\\xi_Q\\}, \\xi_Q\\in Q$\n\\begin{equation}\n\\label{eq:35} \n\\sum\\limits_\\alpha \\f{2^{-l|\\alpha|}}{\\alpha!} \\sum\\limits_Q \n|D_\\xi^\\alpha P_l^\\xi \\sigma(x, \\xi_Q)| \\leq C_n \n2^{ln} \\sup_y \\|P_l^\\xi \\sigma(y, \\cdot)\\|_{L^1({\\mathbf R}^n)}. \n\\end{equation}\nThis is done similar to the $L^2$ case. By \\eqref{eq:859} and by averaging over \nthe corresponding $Q$\n\\begin{eqnarray*}\n& & \\sum\\limits_\\alpha \\f{2^{-l|\\alpha|}}{\\alpha!} \\sum\\limits_Q \n|D_\\eta^\\alpha P_l^\\eta \\sigma(x, \\xi_Q)|\\leq \\sum\\limits_{\\alpha, \\beta} \\f{2^{-l|\\alpha|}}{\\alpha!\\beta!} \n \\sum\\limits_Q |Q|^{-1} \\int_Q |D_\\eta^{\\alpha+\\beta} P_l^\\eta \\sigma(x, \\eta)(\\eta-\\xi_Q)^\\beta|d\\eta\\\\\n & & \\lesssim 2^{ln} \\sum\\limits_{\\alpha, \\beta} \\f{2^{-l|\\alpha|}}{\\alpha!\\beta!} \n \\int |D_\\eta^{\\alpha+\\beta} P_l^\\eta \\sigma(x, \\eta)(\\eta-\\xi_Q)^\\beta|d\\eta\\lesssim 2^{ln} \\|P_l^\\eta \\sigma(x,\n \\cdot)\\|_{L^1} \n\\end{eqnarray*}\n\n\n\n\n\\section{$L^p$ estimates for homogeneous of degree zero symbols}\nWe start with the $L^2$ estimate, since it is very similar to the corresponding \nestimate \\eqref{eq:7} and contains the main ideas for the $L^p$ estimate. \n\n\\subsection{$L^2$ estimates for homogeneous of degree zero symbols}\nConsider $T_\\sigma^*$ and introduce the Littlewood-Paley partition of unity $P_l^{\\xi\/|\\xi|}$. We have \n$$\nT_\\sigma^* g (x)=\\sum\\limits_{l=0}^\\infty \\int e^{2\\pi i \\xi\\cdot x} \n(\\int e^{-2\\pi i \\xi\\cdot y} \n[ g(y) P_l^{\\xi\/|\\xi|} q(y, \\xi\/|\\xi|)] dy )d\\xi\n$$ \nFor every $l\\geq 0$, introduce a partition of unity on $\\mathbf S^{n-1}$, say $\\{K\\}$, which consists of disjoint \nsets of diameter comparable to $2^{-l}$. One may form $\\{ K\\}$ by introducing a $2^{-l}$ net on $\\mathbf S^{n-1}$, say $\\xi^m_{l}$, form the conic sets $H_m^l=\\{\\xi\\in{\\mathbf R}^n: \\: |\\xi\/|\\xi|-\\xi^m_l|\\leq 2^{-l}\\}$ and construct \\\\ $K^l_m=H_m^l\\setminus \\cup_{j=0}^{m-1} H_{j}^l$. We have \n\\begin{equation}\n\\label{eq:400}\nT_\\sigma^* g (x)=\\sum\\limits_{l=0}^\\infty \\sum\\limits_m \\int e^{2\\pi i \\xi\\cdot x} \\chi_{K^l_m}(\\xi)\n(\\int e^{-2\\pi i \\xi\\cdot y} \n[ g(y) P_l^{\\xi\/|\\xi|} q (y, \\xi\/|\\xi|)] dy )d\\xi\n\\end{equation}\nNow, that the symbol is frequency localized around frequencies $\\sim 2^l$ \n and the sets $K^l_m\\cap \\mathbf S^{n-1}$ have diameters less \nthan $2^{-l}$, we expand $q(y, \\xi\/|\\xi|)$ around an {\\it arbitrary point} $\\theta_l^m\\in K_l^m$.\nAccording to \\eqref{eq:36}, we have for all $\\xi\\in K^l_m$, \n$$\nq(y, \\xi\/|\\xi|) \n=\\sum\\limits_{\\alpha\\geq 0} \\f{D_\\xi^\\alpha q(y,\\theta_l^m)}{\\alpha!}(\\xi\/|\\xi|-\\theta_l^m)^\\alpha.\n$$\nEntering this new expression in \\eqref{eq:400} yields \n\\begin{eqnarray*}\n& & \nT_\\sigma^* g (x)=\\sum\\limits_{l=0}^\\infty \\sum\\limits_m \\sum\\limits_{\\alpha} (\\alpha!)^{-1} \n\\int e^{2\\pi i \\xi\\cdot x} \\chi_{K^l_m}(\\xi) (\\xi\/|\\xi|-\\theta_l^m)^\\alpha \\times \\\\\n& & \\times \n(\\int e^{-2\\pi i \\xi\\cdot y} \n[ g(y) D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (y, \\theta_l^m)] dy )d\\xi= \\\\\n& & =\n\\sum\\limits_{l=0}^\\infty \\sum\\limits_m \\sum\\limits_{\\alpha} (\\alpha!)^{-1} Z_{l,m}^\\alpha [g(\\cdot) 2^{-l|\\alpha|} \n D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (\\cdot , \\theta_l^m)] ,\n\\end{eqnarray*}\nwhere $Z_{l,m}^\\alpha$ is given by the multiplier $\\chi_{K^l_m}(\\xi) 2^{l |\\alpha|} \n(\\xi\/|\\xi|-\\theta_l^m)^\\alpha$. Note the disjoint support of the multipliers \n$\\{Z_{l,m}^\\alpha\\}_m$ and $\\|Z_{l, m}^\\alpha\\|_{L^2\\to L^2}=\\sup_\\xi|\\chi_{K^l_m}(\\xi) 2^{l |\\alpha|} \n(\\xi\/|\\xi|-\\theta_l^m)^\\alpha| \\leq 4^{|\\alpha|}$. \\\\\nTake $L^2$ norm of $T_\\sigma^* g$. \n\\begin{eqnarray*}\n& & \\norm{T_\\sigma^* g}{L^2({\\mathbf R}^n)}\\lesssim \n\\sum\\limits_{l=0}^\\infty \\sum\\limits_{\\alpha} (\\alpha!)^{-1}\\left(\\sum\\limits_m \n\\norm{Z_{l, m}^\\alpha [g(\\cdot) 2^{-l|\\alpha|} \n D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (\\cdot , \\theta_l^m) ]}{L^2}^2 \\right)^{1\/2}\\leq \\\\\n& &\\leq 4^{|\\alpha|} \\sum\\limits_{l=0}^\\infty \\sum\\limits_{\\alpha} (\\alpha!)^{-1} \\left(\\sum\\limits_m \\norm{g(\\cdot) 2^{-l|\\alpha|} \n D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (\\cdot , \\theta_l^m) }{L^2}^2\\right)^{1\/2}.\n\\end{eqnarray*}\nWe proceed to further bound the expression in the $m$ sum. Since\n$$\n\\sum\\limits_m \\norm{g(\\cdot) 2^{- l|\\alpha|} \n D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (\\cdot , \\theta_l^m)}{L^2}^2=2^{-2 l|\\alpha|} \\int |g(y)|^2 \\left(\\sum\\limits_m \n| D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (y , \\theta_l^m)|^2\\right) dy, \n$$\nmatters reduce to a good estimate for $\\sum\\limits_m \n| D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (y , \\theta_l^m)|^2$. \nWe proceed as before. By \\eqref{eq:36}, we get for all $\\eta\\in K^l_m\\cap \\mathbf S^{n-1}$, \n\\begin{equation}\n\\label{eq:n3}\n D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(y, \\theta_l^m) \n=\\sum\\limits_{\\beta\\geq 0} \\f{D_\\xi^{\\alpha+\\beta} P_l^{\\eta\/|\\eta|}q(y,\\eta)}{\\beta!}(\\theta_l^m-\\eta)^\\beta.\n\\end{equation}\nAveraging over $K^l_m\\cap \\mathbf S^{n-1}$ and taking into account $|K^l_m\\cap \\mathbf S^{n-1}|\\sim 2^{l(n-1)}$ yields \n\\begin{eqnarray*}\n& & (\\sum\\limits_m \n| D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (y , \\theta_l^m)|^2)^{1\/2} \\lesssim \\\\\n& & \\lesssim \\sum\\limits_{\\beta} \\f{2^{-l |\\beta|}}{\\beta!}\n(\\sum\\limits_m \n|K^l_m\\cap \\mathbf S^{n-1}|^{-1} \\int\\limits_{K^l_m\\cap \\mathbf S^{n-1}} |D_\\xi^{\\alpha+\\beta} P_l^{\\eta\/|\\eta|}q(y,\\eta)|^2\nd\\eta)^{1\/2}\\\\\n& &\\lesssim 2^{l(n-1)\/2} \\sum\\limits_{\\beta} \\f{2^{-l |\\beta|}}{\\beta!} \n\\|D_\\xi^{\\alpha+\\beta} P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^2} \\\\\n& &\\lesssim 2^{l[(n-1)\/2+|\\alpha|]} \n\\sum\\limits_{\\beta} \\f{C_n^{|\\alpha|+|\\beta|}}{\\beta!} \n\\|P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^2(\\mathbf S^{n-1})} \\\\ \n& &\\leq C_n^{|\\alpha|}2^{l[|\\alpha|+(n-1)\/2]}\n\\|P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^2(\\mathbf S^{n-1})}.\n\\end{eqnarray*}\nPutting this back into the estimate for $\\|T_\\sigma^* g\\|_{L^2({\\mathbf R}^n)}$ implies \n$$\n\\norm{T_\\sigma^* g}{L^2({\\mathbf R}^n)}\\lesssim \\|g\\|_{L^2} \\sum\\limits_l 2^{l(n-1)\/2} \\sup\\limits_y \n\\|P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^2(\\mathbf S^{n-1})}\n$$\nas desired. \n\\subsection{$L^p$ estimates for homogeneous of degree zero multipliers}\nFix $p: 2\\leq p<\\leq \\infty$. \nTo verify the estimate $\\|T\\|_{B^0_{p,1}\\to L^p}$, it will suffice to fix $k$ and show \n\\begin{equation}\n\\label{eq:01}\n\\|T(P_k f)\\|_{L^p}\\leq C \\|f\\|_{L^p}.\n\\end{equation}\nFurthermore, by the scale invariance of the quantity $\\sum_l 2^{l(n-1)} \n\\sup_y\\|P_l^{\\xi\/|\\xi|} q(y, \\cdot)\\|_{L^1(\\mathbf S^{n-1})}$ this is equivalent to verifying \\eqref{eq:01} \nonly for $k=0$. \nThat is, it suffices to establish \nthe $L^p, p\\geq 2$ boundedness of the operator \n$$\nG f(x)=\\int\\limits_{{\\mathbf R}^n} q(x,\\xi\/|\\xi|) e^{2\\pi i \\xi x} \\varphi(|\\xi|) \\hat{f}(\\xi) d\\xi.\n$$\nprovided the multiplier $m$ satisfies \n$\\sum\\limits_l 2^{l(n-1)} \\sup_y\\|P_l^{\\xi\/|\\xi|} q(y, \\cdot)\\|_{L^1(\\mathbf S^{n-1})}<\\infty$. \n\nNext, we make the angular decomposition as in the case of the $L^2$ estimates for the adjoint \noperator $G^*$.However, this time we will have to be more careful and \ninstead of the rough cutoffs $\\chi_{K^l_m}$, we shall use a \nsmoothed out versions of them. Fix $l$. Choose and fix a \n $2^{-l}$ net $\\theta_m^l\\in K_m^l\\cap \\mathbf S^{n-1}$, so that the family \n$\\{\\theta\\in \\mathbf S^{n-1}: \n|\\theta_m^l-\\theta|\\leq 2^{-l}\\}_m$ has the finite intersection\nproperty. \nIntroduce a family of functions $\\varphi_{l,m}:{\\mathbf R}^n\\to [0,1]$, so that for every $\\xi\\in {\\mathbf R}^n$, \n\\begin{eqnarray}\n\\label{eq:fun}\n& & \\sum\\limits_m \\varphi_{l,m} (2^l(\\xi\/|\\xi|-\\theta_m^l))=1 \\\\ \n\\nonumber\n& & \\sup_\\eta |D^\\beta_\\eta \\varphi_{l,m} (\\eta)|\\leq C_\\beta.\n\\end{eqnarray}\nIn other words, the family of functions $\\{\\varphi_{l,m}\\}$ provides a \nsmooth partition of unity, subordinated to the cover $\\{K_m^l\\}$. \n\nAs before, write \n$$\nG^* g(x)=\\sum\\limits_{l\\geq 0} \\int\\limits_{{\\mathbf R}^n} e^{2\\pi i \\xi x} \\varphi(|\\xi|) \\int e^{-2\\pi i \\xi y} [g(y) \nP_l^{\\xi\/|\\xi|} q(y, \\xi\/|\\xi|)]dy d\\xi.\n$$\nInserting the partition of unity discussed above into the ($l^{th}$ term of the) \nlast formula for $G^*$ yields \n$$\nG^* g (x)=\\sum\\limits_{l\\geq 0} \\sum\\limits_m \\int e^{2\\pi i \\xi (x-y)} \n\\varphi_{l,m} (2^l(\\xi\/|\\xi|-\\theta_m^l)) \\varphi(|\\xi|) [g(y) \nP_l^{\\xi\/|\\xi|} q(y, \\xi\/|\\xi|)]dy d\\xi. \n$$\nFollowing the same strategy as before, we expand $q(y, \\xi\/|\\xi|)$ around $\\theta_m^l\\in K_m^l$.\nAccording to \\eqref{eq:36}, we have \n$$\nP_l^{\\xi\/|\\xi|} q(y, \\xi\/|\\xi|) \n=\\sum\\limits_{\\alpha\\geq 0} \\f{D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} \nq(y, \\theta_m^l) }{\\alpha!}(\\xi\/|\\xi|-\\theta_m^l)^\\alpha.\n$$\nOf course, the last formula is useful only when \n$|\\xi\/|\\xi|-\\theta_m^l|\\lesssim 2^{-l}$, in particular on the \nsupport of $\\varphi_{l,m} (2^l(\\xi\/|\\xi|-\\xi_m^l)) $. \nThis gives us the representation \n\\begin{eqnarray*}\n& & G^* g =\\sum\\limits_{l\\geq 0}\\sum\\limits_m \\sum\\limits_{|\\alpha|\\geq 0} (\\alpha!)^{-1}\n\\int e^{2\\pi i \\xi x} \n\\varphi_{l,m} (2^l(\\xi\/|\\xi|-\\theta_m^l)) (\\xi\/|\\xi|-\\theta_m^l)^\\alpha \\varphi(|\\xi|) \\times \\\\\n& &\\times \\int e^{2\\pi i \\xi y} g(y) \nP_l^{\\xi\/|\\xi|} D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(y, \\theta_m^l)dy d\\xi= \\\\\n& & =\n\\sum\\limits_{l\\geq 0}\\sum\\limits_m \\sum\\limits_{|\\alpha|\\geq 0}(\\alpha!)^{-1} Z_{l,m}^\\alpha \n[g(\\cdot) 2^{-l|\\alpha|} D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(\\cdot, \\theta_m^l)]\n\\end{eqnarray*}\nwhere \n$$\n\\widehat{Z_{l,m}^\\alpha f}(\\xi)= \\varphi_{l,m} (2^l(\\xi\/|\\xi|-\\theta_m^l)) 2^{l|\\alpha|} \n(\\xi\/|\\xi|-\\theta_m^l)^\\alpha \n\\varphi(|\\xi|) \\hat{f}(\\xi)=\\varphi_{l,m}^\\alpha (\\xi\/|\\xi|-\\theta_m^l)\\varphi(|\\xi|) \\hat{f}(\\xi).\n$$\nTaking $L^p$ norm of $G^* g$, we get \n\\begin{eqnarray*}\n& & \\|G^* g\\|_{L^p}\\leq \\sum\\limits_{l\\geq 0} \\sum\\limits_{|\\alpha|\\geq 0}(\\alpha!)^{-1}\\|\\sum\\limits_m Z_{l,m}^\\alpha \n[g(\\cdot) 2^{-l|\\alpha|} D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(\\cdot, \\theta_m^l)]\\|_{L^p}\n\\end{eqnarray*}\nLemma \\ref{le:sum} in the Appendix allows us to treat expressions of the type \n$\\|\\sum\\limits_m Z_{l,m}^\\alpha g_m^\\alpha\\|_{L^p}$. Indeed, according to \\eqref{eq:n1}, we have \n\\begin{eqnarray*}\n& & \\|\\sum\\limits_m Z_{l,m}^\\alpha \n[g(\\cdot) 2^{-l|\\alpha|} D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(\\cdot, \\theta_m^l)]\\|_{L^p}\\lesssim (\\sum\\limits_m \n\\|g(\\cdot) 2^{-l|\\alpha|} D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(\\cdot, \\theta_m^l)\\|_{L^p}^p)^{1\/p} \\\\\n& &= 2^{-l|\\alpha|} (\\int |g(y)|^p (\\sum\\limits_m \n |D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(y, \\theta_m^l)|^p) dy)^{1\/p} \n\\end{eqnarray*}\nBy virtue of \\eqref{eq:n3}, we get \n\\begin{eqnarray*}\n& & \nD_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q(y, \\theta_l^m) \n=\\sum\\limits_{\\beta\\geq 0} \\f{D_\\xi^{\\alpha+\\beta} P_l^{\\eta\/|\\eta|}q(y,\\eta)}{\\beta!}(\\theta_l^m-\\eta)^\\beta.\n\\end{eqnarray*}\nwhence by averaging\\footnote{this step is identical to the one performed earlier for the \n$L^2$ bounds, except that now the $l^2$ sums are replaced by $l^p$ sums.} over $K^l_m\\cap \\mathbf S^{n-1}$, \n\\begin{eqnarray*}\n & & (\\sum\\limits_m \n| D_\\xi^\\alpha P_l^{\\xi\/|\\xi|} q (y , \\theta_l^m)|^p)^{1\/p} \\lesssim \\\\\n& & \\lesssim \\sum\\limits_{\\beta} \\f{2^{-l |\\beta|}}{\\beta!}\n(\\sum\\limits_m \n|K^l_m\\cap \\mathbf S^{n-1}|^{-1} \\int\\limits_{K^l_m\\cap \\mathbf S^{n-1}} |D_\\xi^{\\alpha+\\beta} P_l^{\\eta\/|\\eta|}q(y,\\eta)|^p\nd\\eta)^{1\/p}\\\\\n& &\\lesssim 2^{l(n-1)\/p} \\sum\\limits_{\\beta} \\f{2^{-l |\\beta|}}{\\beta!} \n\\|D_\\xi^{\\alpha+\\beta} P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^p(\\mathbf S^{n-1})} \\\\\n& &\\lesssim 2^{l[(n-1)\/p+|\\alpha|]} \n\\sum\\limits_{\\beta} \\f{C_n^{|\\alpha|+|\\beta|}}{\\beta!} \n\\|P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^p(\\mathbf S^{n-1})} \\\\ \n& &\\leq C_n^{|\\alpha|}2^{l[|\\alpha|+(n-1)\/p]}\n\\|P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^p(\\mathbf S^{n-1})}.\n\\end{eqnarray*}\nAll in all, \n\\begin{eqnarray*}\n& & \\|G^* g\\|_{L^p}\\leq C_n \\|g\\|_{L^p}\\sum\\limits_{l\\geq 0} \\sum\\limits_{|\\alpha|\\geq 0}(\\alpha!)^{-1} \n C_n^{|\\alpha|}2^{l(n-1)\/p}\n\\sup\\limits_y \\|P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^p(\\mathbf S^{n-1})}\\\\ \n& &\\leq C_n \\|g\\|_{L^p} \\sum\\limits_l \n2^{l(n-1)\/p}\n\\sup\\limits_y \\|P_l^{\\eta\/|\\eta|}q(y,\\cdot)\\|_{L^p(\\mathbf S^{n-1})}.\n\\end{eqnarray*}\nas desired. \n\n\\section{Counterexamples}\n\\label{sec:counter}\n\\subsection{Theorem \\ref{theo:1} is sharp.}\nGiven $p>2$, we will construct \n an explicit symbol $\\sigma(x, \\xi)$, \nso that the corresponding PDO $T_\\sigma$ is not bounded on $L^2({\\mathbf R}^n)$, \nbut which satisfies \\\\\n$\\sup_x |D_\\xi^\\alpha \\sigma(x, \\xi)|\\leq C_\\alpha |\\xi|^{-|\\alpha|}$ and \n$\\sup_x \\norm{\\sigma(x, \\cdot)}{W^{p,n\/p}}<\\infty$. \nThe construction is a minor modification of the \nstandard example of a symbol in $S^0_{1, 1}$, which \nis not bounded on $L^2$, see for example \\cite{Stein}, page 272. \nWe carry out the construction in $n=1$, \nbut this can be easily generalized to higher dimensions.\n\nFor the given $p>2$, fix small $0<\\delta<1\/2$, \nso that\\footnote{The reason for this choice of $\\delta$ will become apparent in the proof below.}\n $2+4\\delta\/(1-2\\delta) 1$, \n$$\n\\sup_x |D_\\xi^\\alpha \\sigma(x, \\xi)|\\sim |\\xi|^{-|\\alpha|} \\ln^{\\delta-1\/2}(|\\xi|) \n\\leq |\\xi|^{-|\\alpha|}.\n$$\nFinally, to \nestimate $\\sup_x \\norm{\\sigma(x, \\cdot)}{W^{p,1\/p}}$, write \n$$\n\\sigma=\\sum\\limits_{s=3}^\\infty \\sum\\limits_{j=2^s}^{2^{s+1}} \\f{e^{-2\\pi i 2^j x}}{j^{1\/2-\\delta}}\n \\varphi(2^{-j} \\xi)=:\\sum\\limits_{s=3}^\\infty \\sigma^s,\n$$\nBy the convexity of the norms, we have with $\\theta: 1\/p=\\theta\/2$, \n$$\n\\norm{\\sigma^s(x, \\cdot)}{W^{p,1\/p}}\\leq \n\\norm{\\sigma^s(x, \\cdot)}{H^{1\/2}}^\\theta\n\\norm{\\sigma^s(x, \\cdot)}{L^\\infty}^{(1-\\theta)}\n$$\nIt is now easy to compute the norms on the right hand side. We have \n$$\n\\sup_x \\norm{\\sigma^s(x, \\cdot)}{H^{1\/2}} \n\\sim (\\sum_{j=2^{s}}^{2^{s+1}} \\f{1}{j^{1-2\\delta}})^{1\/2}\n\\sim 2^{\\delta s}.\n$$\nOn the other hand, \n$$\n\\norm{\\sigma^s(x, \\cdot)}{L^\\infty}\\sim 2^{-s(1\/2-\\delta)},\n$$\nwhence $\\sup_x \\norm{\\sigma^s(x, \\cdot)}{W^{p,1\/p}}\\leq \n2^{s(\\delta\\theta-(1\/2-\\delta)(1-\\theta))}$. Clearly, \nsuch an expression dyadically sums in $s\\geq 3$, \nprovided $\\delta\\theta<(1\/2-\\delta)(1-\\theta)$ or equivalently $p>2+4\\delta\/(1-2\\delta)$.\n\\subsection{Proposition \\ref{prop:5}: Theorem \\ref{theo:4} is sharp}\n\\begin{proof}(Proposition \\ref{prop:5}) \nWe construct a sequence of symbols $\\sigma_\\delta:\\mathbf R^2\\times \\mathbf R^2\\to\\mathbf R^1$, \nso that for a fixed Schwartz function $f$\n$$\n\\lim_{\\delta\\to 0+} |T_{\\sigma_\\delta} f|=|H_* f(x)|=\\sup_{u\\in \\mathbf S^1} |H_u f(x)|\n$$\nSince we already know, \\cite{LL}, that \n$H_*$ is {\\it unbounded} on $L^2(\\mathbf R^2)$, \nwe should have\n\\begin{equation}\n\\label{eq:715}\n\\limsup_{\\delta\\to 0+} \\|T_{\\sigma_\\delta}\\|_{L^2\\to L^2}=\\infty. \n\\end{equation}\nIn our construction \n$\\sigma_\\delta$ will depend on $f$, but it is still clear \nthat one can achieve \\eqref{eq:715}. \nNamely, take a sequence $f_N: \\norm{f_N}{L^2(\\mathbf R^2)}=1$, \nso that $\\|H_* (f_N)\\|_{L^2(\\mathbf R^2)}\\geq N$. \nThen construct $\\sigma_{N, \\delta}$, so that $\\lim_{\\delta\\to 0+} \n|T_{\\sigma_{\\delta, N}} f_N|=\nH_* f_N$. Then clearly, \\\\\n$\\limsup_{ N\\to \\infty, \\delta\\to 0+} \n\\|T_{\\sigma_{N,\\delta}}\\|_{L^2\\to L^2}=\\infty$. \n\nNow, from the $L^2$ boundedness results of \nTheorem \\ref{theo:4} (or rather the lack thereof), we must have \n\\begin{equation}\n\\label{eq:713}\n\\limsup_{\\delta\\to 0} \\sum\\limits_l 2^{l\/2} \n\\sup_x \\|P_l^{\\xi\/|\\xi|} \\sigma_\\delta(x, \\cdot)\\|_{L^2(\\mathbf S^1)}=\\infty.\n\\end{equation}\nOn the other hand, we will see that \n$\\sup_{x, \\xi, \\delta}|\\sigma_\\delta(x, \\xi)|\\leq 1$ and \n\\begin{equation}\n\\label{eq:714}\n\\sup_{\\delta, x} \\|\\sigma_\\delta(x, \\cdot)\\|_{W^{1, 1}(\\mathbf S^1)}<\\infty. \n\\end{equation}\nNote in contrast that (at least heuristically) \\eqref{eq:713} states \n$$\n\\limsup_{\\delta\\to 0} \\sup_x \\|\\sigma_\\delta(x, \\cdot)\\|_{B^{1\/2}_{2, 1}}=\\infty\n$$\nand by the Sobolev embedding estimate on the sphere \n\\eqref{eq:bern} (and up to the usual Besov spaces \nadjustments at the endpoints), one should have that the quantity \nin \\eqref{eq:714} (at least in principle) controls \\eqref{eq:713}. Having both \n\\eqref{eq:713} and \\eqref{eq:714} for a concrete example suggests that \nthe conditions imposed in Theorem \\ref{theo:4} \n are extremely tight. \n\nLet us now describe the construction of $\\sigma_\\delta$. First of all, \n\\begin{eqnarray*}\nH_*f(x) = \\sup_{u\\in\\mathbf S^1} |H_u f(x)| &=& \\sup_{u\\in\\mathbf S^1}|\\int \nsgn(u\\cdot \\xi\/|\\xi|) \\hat{f}(\\xi) e^{2\\pi i \\xi\\cdot x} d\\xi|=\\\\\n&=&\n|\\int \nsgn(u(x)\\cdot \\xi\/|\\xi|) \\hat{f}(\\xi) e^{2\\pi i \\xi\\cdot x} d\\xi|,\n\\end{eqnarray*}\nfor some measurable function $u(x):\\mathbf R^1\\to \\mathbf S^1$. \nClearly $u(x)$ will depend on the function $f$, see the remarks above after \\eqref{eq:715}. \\\\\nIntroduce a function $\\psi:\\psi\\in C^\\infty, -1\\leq \\psi(x)\\leq 1$, and \nso that $\\psi(z)=-1: z\\in (-\\infty, -1]$, $\\psi(z)=1: \nz\\in [1, \\infty)$. Clearly \n$$\nH_* f(x)=\\lim_{\\delta\\to 0+} T_{\\sigma_\\delta} f(x)= \\lim_{\\delta\\to 0+}|\\int \n\\psi\\left(\\f{u(x)\\cdot \\xi\/|\\xi|}{\\delta}\\right) \n\\hat{f}(\\xi) e^{2\\pi i \\xi\\cdot x} d\\xi|, \n$$\nthat is $\\sigma_\\delta(x, \\xi\/|\\xi|)=\\psi(\\f{u(x)\\cdot \\xi\/|\\xi|}{\\delta})$, \nfor which we will verify \\eqref{eq:714}, while it \nis clearly bounded in absolute value by one. \n\nWe pause \nfor a second to comment on the particular form of $T_{\\sigma_\\delta}$. Note that the function $u(x)$ \nin general will not be smooth\\footnote{Note that under some extra smoothness assumptions on $u$, \nLacey and Li have managed to prove $L^2$ boundedness!} and therefore will not fall under \nthe scope of any standard boundedness theory for PDO. Also, note that while the map \n $\\xi\\to \\sigma_\\delta(x, \\xi)$ is definitely smooth, its derivatives are quite large and \nblow up at the important limit $\\delta\\to 0$. This shows that in order to \ntreat maximal operators, build upon singular multipliers (as is the case here), \none needs the full strength of Theorems \\ref{theo:1}, \\ref{theo:4} and beyond. \n\nGoing back to the proof of \\eqref{eq:714}, compute \n\\begin{eqnarray*}\n& & \\f{\\partial \\sigma}{\\partial_{\\xi_1}}= \n\\f{\\psi'(\\f{u(x)\\cdot \\xi\/|\\xi|}{\\delta})}{\\delta|\\xi|^3}\\left(u_1(x)\\xi_2^2- u_2(x)\\xi_1\\xi_2\\right)= \n\\f{\\psi'(\\f{u(x)\\cdot \\xi\/|\\xi|}{\\delta})\\xi_2}{\\delta|\\xi|^2}u(x)\\cdot \n(\\xi\/|\\xi|)^{\\perp} \\\\\n& & \\f{\\partial \\sigma}{\\partial_{\\xi_2}}=\n\\f{\\psi'(\\f{u(x)\\cdot \\xi\/|\\xi|}{\\delta})}{\\delta|\\xi|^3}\n\\left(u_1(x)\\xi_1^2- u_2(x)\\xi_1\\xi_2\\right)= \\f{\\psi'(\\f{u(x)\\cdot \n\\xi\/|\\xi|}{\\delta})\\xi_1}{\\delta|\\xi|^2} u(x)\\cdot \n(\\xi\/|\\xi|)^{\\perp} \n\\end{eqnarray*}\nClearly, the supports of both derivatives are in \n$\\xi: |u(x)\\cdot \\xi\/|\\xi||\\leq \\delta<<1$. Also, on their support, \n$|\\nabla \\sigma(x, \\xi\/|\\xi|)|\\sim C \\delta^{-1}$. It follows \n$$\n\\|\\sigma_\\delta(x, \\cdot)\\|_{W^{1, 1}(\\mathbf S^1)}\\leq \n\\int_{\\xi\\in \\mathbf S^1: |u(x)\\cdot \\xi\/|\\xi|\\leq \\delta}\n|\\nabla \\sigma_\\delta(x, \\xi)|d\\xi\\leq C,\n$$\nwhere $C$ is independent of $\\delta$. This was the claim in \\eqref{eq:714}. \n\\end{proof}\n\n\n\n\n\n\\section{Appendix}\n\n\\subsection{Estimates for Fourier transforms of functions supported on small spherical caps.}\nIn this section, we present a pointwise estimate for the kernels of \nmultipliers that restrict the Fourier transform to a small spherical cap. \n\\begin{lemma}\n\\label{le:900}\nLet $\\theta\\in \\mathbf S^{n-1}$ and \n$\\varphi$ is a $C^\\infty$ function with $supp \\ \\varphi\\subset \\{\\xi: 1\/2\\leq |\\xi|\\leq 2\\}$. \nLet also $l>0$ be any integer. Define $K_{l, \\theta}$ to be the inverse \nFourier transform of $\\varphi(2^l(\\xi\/|\\xi|-\\theta))\\varphi(|\\xi|)$, that is \n$$\nK_{l, \\theta} (x)=\\int \\varphi(2^l(\\xi\/|\\xi|-\\theta))\\varphi(|\\xi|) e^{2\\pi i x\\cdot \\xi} d\\xi.\n$$\nThen, for every $N>0$, there exists $C_N$, so that \n\\begin{equation}\n\\label{eq:201}\n|K_{l, \\theta}(x)|\\leq C_N 2^{-l(n-1)}(1+|\\dpr{x}{\\theta}|)^{-N} \n(1+2^{-l}|x-\\dpr{x}{\\theta}\\theta|)^{-N}.\n\\end{equation}\nThat is, in the direction of $\\theta$, the function has any polynomial decay, while in the directions transversal to $\\theta$, one has decay like $(2^{-l}