diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzprcg" "b/data_all_eng_slimpj/shuffled/split2/finalzzprcg" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzprcg" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction} \nControlling the photonic transport in integrated circuits is important for building future information and communication technologies. While a remarkable progress has been achieved on early implementations of such technologies \\cite{Asavanant,Larsen,Juan}, the development of an optical rectifier that can act on single or a few photon pulses remain challenging. These rectifying devices are useful for routing and isolating signals. Rectifiers based on ferromagnetic compounds have been demonstrated for increasing signal processing capabilities. However, they are inherently lossy and difficult to miniaturize and use in integrated circuits \\cite{Pozar}. Recently, using a semi-classical theory, a two atoms device has been identified as a candidate for exhibiting strong directionality \\cite{auffeves2014}. The transmission of light through this device would depend on the direction of propagation. If the input comes from one side it will pass through it, and if it comes from the other side it will get reflected. This has triggered a debate over whether such a device can perform with non-classical states of light \\cite{Alex,Auffeves2016,Vienna2016,Combes,Fang,Clemens,Gonzalez,Ordonez}.\n\nThis rectifying device, also referred to as \\textit{optical diode}, is consisting of a pair of two-level systems coupled to a one-dimensional waveguide. By changing the distance between the atoms and the detuning of their resonance frequencies, the efficiency of the diode was optimized. It has been shown that for light pulses in coherent states, the device can exhibit efficiencies of more than 60\\% \\cite{Alex,Auffeves2016,Vienna2016,Combes}. However, in the monochromatic limit, it has been found that this device cannot rectify single photon pulses \\cite{Alex}. Since significant directionality can be achieved for coherent inputs while a symmetric behaviour is predicted for single photon pulses, one might naturally ask: which components of the coherent state triggers the performance of the device? Do superpositions matter? How can we change these critical components, so we can improve the efficiency of this device for inputs with a few photon pulses? To answer this questions, we will use the same approach as in \\cite{Alex}. Using the Heisenberg equations of motion, we will show that by changing the length of the light pulse, we can reduce the number of photons necessary to activate the rectifying behaviour of the optical diode. \n\nThis article is organized as follow: in section II, we will introduce the theoretical model for this device and present how the rectification factor of the device is computed. In section III, we will discuss the cases of photon-number pulses. We will show that the main terms that will contribute to the rectification depend directly on the pulse duration. In section IV, we will present a few concluding remarks and discuss potential future directions of this research.\n\n\\section{The model}\n\n\\subsection{The rectifying device}\n\nAs illustrated in Fig.~\\ref{fig:diode}, the system of interest here is composed of a pair of atoms strongly coupled to a one dimensional waveguide. Each atom is model as a two-level system, for which $\\ket{g_j}$ and $\\ket{e_j}$ are the ground and excited states of the $j^{th}$ atom, respectively. The atom on the right is resonant with the central frequency of the light pulse $\\omega_0$. On the other hand, the atom on the left is detuned. We denote its resonance frequency by $\\omega_1$. Both frequencies are assumed to be much larger than the cutoff frequency of the waveguide \\cite{Snyder}. In this case, the dispersion relations read $\\omega = v_g \\abs{k}$, such that $k$ is the longitudinal wave-number of the field mode and $v_g$ is the group velocity of the light pulse in the waveguide. The free Hamiltonian of this system reads\n\\begin{equation} \\label{H0}\n \\hat{\\mathcal{H}}_{0} = \\sum_{j=1}^{2} \\hbar \\omega_j \\ketbra{e_j} + \\int_{0}^{\\infty} \\dd \\omega \\hbar \\omega (\\hat{a}^{\\dagger}_{\\omega}\\hat{a}^{}_{\\omega} + \\hat{b}^{\\dagger}_{\\omega}\\hat{b}^{}_{\\omega}) \\;,\n\\end{equation}\nsuch that the first term corresponds to the two atoms. The second term describes the field modes. The operator $\\hat{a}^{\\dagger}_{\\omega}$ ($\\hat{b}^{\\dagger}_{\\omega}$) creates a photon with frequency $\\omega$ moving towards the right (left). The interaction between the atoms and the propagating light pulse is described by the dipole Hamiltonian within the rotating wave approximation \\cite{Snyder,Roulet}\n\\begin{equation} \\label{Hdip}\n\\begin{split}\n \\hat{\\mathcal{H}}_{\\mathrm{dip}} =& -i \\hbar \\sum_{j=1}^{2} \\int_{0}^{\\infty} \\dd \\omega g_{\\omega}^{(j)} \\\\\n \\times \\Big[ & \\hat{\\sigma}_{+}^{(j)} \\left( \\hat{a}_{\\omega} e^{i \\omega \\frac{ x_j}{v_g}} + \\hat{b}_{\\omega} e^{- i \\omega \\frac{ x_j}{v_g}}\\right)e^{ - i (\\omega - \\omega_j)t} - \\textbf{H.c.} \\Big] \\;.\n\\end{split}\n\\end{equation}\nHere $x_j$, $g_{\\omega}^{(j)}$ and $\\hat{\\sigma}_{+}^{(j)} = \\ketbra{e_j}{g_j} $ are the position, the coupling strength and raising operator of the $j^{th}$ atom, respectively. Similarly to the previous study \\cite{Alex}, we assume the Weisskopf-Wigner approximation \\cite{Scully}, i.e. both atoms have the same coupling to the waveguide, $g_{\\omega}^{(1)} = g_{\\omega}^{(2)} = g$. Thus, the decay rate to the waveguide for both atoms is $\\gamma = 2\\pi g^2$.\n\n\\subsection{Computing rectification}\n\nSeveral figures of merit have been used in previous papers to capture rectification: we review them in Appendix \\ref{app:defs}. Here we define the rectifying factor of the device as\n\\begin{equation}\\label{DiodeEff}\n \\mathcal{R} = T_{\\rightarrow} - T_{\\leftarrow}\\,,\n\\end{equation} where $T_{\\rightarrow}$ is the transmittivity of the light coming from the left (i.e.~the fraction of left-incoming light scattered towards the right); analogously, $T_{\\leftarrow}$ is the fraction of right-incoming light scattered towards the left. Clearly $-1\\leq\\mathcal{R}\\leq 1$, the sign determining the direction in which the diode rectifies. Explicitly,\n\\begin{eqnarray}\\label{eqsT}\nT_{\\rightarrow}&=&\\frac{1}{F}\\lim_{t\\rightarrow\\infty}N_a^{\\rightarrow}(t)\\,=\\,\\frac{1}{F}(1-\\lim_{t\\rightarrow\\infty}N_b^{\\rightarrow}(t))\\,,\\\\\nT_{\\leftarrow}&=&\\frac{1}{F}\\lim_{t\\rightarrow\\infty}N_b^{\\leftarrow}(t)\n\\end{eqnarray}\nwhere $F$ is the mean photon flux per unit time, and where \\begin{eqnarray*}\nN_{b}^{\\textrm{input}} (t) = \\int_{0}^{\\infty} \\dd \\omega \\textrm{Tr}\\left[\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}^{}_{\\omega}(t)\\,\\rho_{\\textrm{input}}\\right]\\;,\n\\end{eqnarray*} (and similar definitions for similar quantities).\n\nIn this paper, we shall consider monomode Fourier-limited pulses. The mode that defines a pulse coming from the left is \n\\begin{equation}\\label{hatA}\n\\hat{A}^\\dagger = \\int_{0}^{\\infty} \\dd\\omega f (\\omega) \\hat{a}_{\\omega}^{\\dagger} = \\int_{0}^{\\infty} \\dd\\tau \\xi (\\tau) \\hat{a}_{\\tau}^{\\dagger} \\;,\n\\end{equation}\nwhere\n\\begin{equation}\n \\begin{split}\n \\hat{a}_{\\tau}^{} =& \\frac{1}{\\sqrt{2\\pi}} \\int \\dd\\omega \\hat{a}_{\\omega}^{} e^{- i (\\omega - \\omega_0) \\tau} \\;,\\\\\n \\xi (\\tau) =& \\frac{1}{\\sqrt{2\\pi}} \\int \\dd\\omega f (\\omega) e^{- i (\\omega - \\omega_0) \\tau}\\,.\n \\end{split}\n\\end{equation} For pulses coming from the right, we use the identical definition of a mode $\\hat{B}$, replacing the operators $(\\hat{a}_{.},\\hat{a}^\\dagger_{.})$ with $(\\hat{b}_{.},\\hat{b}^\\dagger_{.})$.\n\nLike in previous works, for definiteness the calculations will assume a square pulse\n\\begin{equation}\\label{xitau}\n\\xi (\\tau) = \\Bigg\\{\n \\begin{array}{cl}\n \\sqrt{\\Omega\/2} & \\text{for } 0 \\leq \\tau \\leq 2\/\\Omega, \\\\\n 0 & \\text{otherwise}.\n \\end{array}\n\\end{equation}\nFor an input state $\\ket{\\psi}$, the mean flux of photons coming from the left is \n\\begin{equation}\\label{flux}\n F = \\textrm{Tr}\\left[\\hat{a}_\\tau^\\dagger(0)\\hat{a}_\\tau(0)\\rho_{\\textrm{input}}\\right]\\;,\n\\end{equation}\n\nA basis of the total states of the system is denoted $\\ket{n_a,n_b,s_1,s_2}$ where $n_{a,b}$ is a Fock state in mode $\\hat{A}$ or $\\hat{B}$, while $s_j\\in\\{g,e\\}$ is the state of the $j$-th atom. Initially, the pulse will come from either the left or the right; and both atoms are assumed to be in the ground state to ensure that the rectifying device is passive. Thus our input states will always be of the form $\\ket{\\psi_a}\\equiv\\ket{\\psi,0,g,g}$, or $\\ket{\\psi_b}\\equiv\\ket{0,\\psi,g,g}$.\n\nInstead of solving the time-dependent equations of motion and sending $t\\rightarrow\\infty$, we shall consider only long pulses and thus assume that the system is in the steady state:\n\\begin{eqnarray}\n\\lim_{t\\rightarrow\\infty}N_b^{\\textrm{input}}(t)&\\approx&\\int_{0}^{\\infty} \\dd \\omega \\expval{\\hat{b}^{\\dagger}_{\\omega}\\hat{b}^{}_{\\omega}}_{ss,\\textrm{input}}\\,.\\label{ss:assume}\n\\end{eqnarray}\n\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=0.5\\textwidth,trim={0 9cm 0 0cm},clip]{figures\/Diode.pdf}\n \\caption{A scheme of the rectifying device consisting of two 2-level systems coupled to a one-dimensional waveguide. The one on the right is resonant with the central frequency of the light pulse, $\\omega_0$. The atom on the left is detuned, $\\omega_1 \\neq \\omega_0$.}\n \\label{fig:diode}\n\\end{figure}\n\n\\subsection{The case for pulses of finite length}\n\nAll previous fully-quantum studies of this rectifying device \\cite{Alex,Vienna2016,Auffeves2016} concluded that there is no rectification for single-photon Fock states. This conclusion was drawn in the limit of pulses of \\textit{infinite length}. One of the results of the next Section will be that \\textit{the device does show some rectification for single-photon Fock states} for pulses of finite length.\n\nThose same works also showed that high rectification is expected for coherent states, in qualitative (if not exact quantitative) agreement with a previous semi-classical study \\cite{auffeves2014}. Specifically, rectification was predicted \\cite{Alex} and later observed \\cite{Clemens} to have a high-value plateau when $F\/\\gamma$ ranges between $10^{-3}$ and $10^{-1}$. Let us now describe this case in more detail, to understand what may be the origin of the rectification.\n\nEvery coherent state is ultimately monomode \\cite{Alex90s,Alex90ss}: a coherent state pulse with average number of photons $\\bar{n}$, propagating towards the right, can be written as\n\\begin{equation}\\label{state:cohbis}\n \\ket{\\alpha_a} = \\exp\\big(\\sqrt{\\bar{n}} (\\hat{A}^\\dagger - \\hat{A})\\big) \\ket{\\text{\\o{}}}\\;.\n\\end{equation} Since $\\hat{a}_\\tau(0)\\ket{\\alpha} = \\sqrt{\\bar{n}}\\, \\xi(\\tau) \\ket{\\alpha}$, the flux \\eqref{flux} is given by $F=\\frac{\\bar{n} \\Omega}{2}$. If the device performance is maximized for plateau centred around $F\/\\gamma \\approx 10^{-2}$, then the optimal average number of photons in the pulse is\n\\begin{equation}\n \\bar{n}_{\\textrm{opt}} \\approx \\frac{2 \\gamma }{\\Omega} \\times 10^{-2}\\,.\n\\end{equation}\nIf we were to consider infinitely long pulses, as used in the single-photon case, then we would have to take the monochromatic limit $\\Omega \\rightarrow 0$; in which case $\\bar{n}_{\\textrm{opt}}$ diverges.\n\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=0.23\\textwidth]{figures\/f1s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p1pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f2s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p2pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f3s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p3pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f4s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p4pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f5s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p5pow.pdf}\\\\\n \n \\includegraphics[width=0.23\\textwidth]{figures\/barre.pdf}\n \\caption{The diode rectification $\\mathcal{R}$ for different Fock states ($n=1,2,3,4$ and 5 photons, respectively from top to bottom). Left panels: the rectification factors are computed as a function of the detuning $\\Delta = \\omega_0 - \\omega_1$, and the phase $\\theta = \\omega_0 d\/v_g$ while the bandwidth of the input pulse is fixed, ($\\gamma\/\\Omega = 10^{2}$. On the right panels, we plot the rectification factors for a fixed detuning, $\\Delta\/\\gamma = 0.1$ while we vary the bandwidth $\\Omega$.}\n \\label{fig:eff}\n \\vspace{-2cm}\n\\end{figure}\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=0.48\\textwidth]{figures\/RectDelta.pdf}\n \\vspace{-2mm}\n \\caption{The diode rectification $\\mathcal{R}$ for different Fock states ($n=1,2,3,4, 5$ and 22 photons). The pulse length is fixed ($\\Omega\/\\gamma = 10^{-2}$). The rectification factors are computed as a function of the detuning $\\Delta = \\omega_0 - \\omega_1$. For the case of $n=1$, the phase $\\theta$ is optimized in order to have the strongest negativity of the rectification factor. Ergo, we set $\\theta \/ 2\\pi = 0.5025$. Then we used the same value for the other cases. Here, we show the case of $n=22$ because for this Fock state we observed the strongest negative rectification.}\n \\label{fig:negeff}\n\\end{figure}\n\n\\section{Beyond the monochromatic case}\n\nIn this section, we consider the case of a pulse with a finite length. However, for our approximation \\eqref{ss:assume} to be valid, we need the pulse to be long enough such that the transient regime remains negligible. We shall set the bandwidth to be two orders of magnitude smaller than the decay rate to the waveguide, $\\Omega\/\\gamma = 10^{-2}$. For such a value, the necessary average number of photons to be in the middle of the plateau for the coherent state should be $\\bar{n}_{\\textrm{opt}} \\approx 2$.\n\n\\subsection{The device is not sensitive to coherences between Fock states}\n\nWe first prove that coherences between Fock states don't play any role in the dynamics of the rectification device. Consider as input state a generic superposition\n\\begin{eqnarray}\n\\ket{\\psi}&=&\\sum_{n=0}^\\infty c_n\\ket{n} \\;,\n\\end{eqnarray} \nwith $\\ket{n}$ standing for the Fock state with $n$ photons. For definiteness, we look at a special case, but the argument is the same for any of the calculations of interest here.\n\nWe assume therefore that the state $\\ket{\\psi}$ is prepared in mode $\\hat{A}$, i.e.~the input state is $\\ket{\\psi_a}$ with the Fock states given by \\cite{Alex90s,Alex90ss}\n\\begin{equation}\\label{state:Fockn}\n \\ket{n} = \\frac{(\\hat{A}^{\\dagger})^n}{\\sqrt{n!}} \\ket{\\text{\\o{}}}\\,.\n\\end{equation} Then we look at the number of reflected photons\n\\begin{equation}\\label{eq:NnmAll}\n N_{b}^{\\psi} (t) = \n \\sum_{m,n}c_nc_m^*\\int_{0}^{\\infty}\\dd\\omega\\bra{m_a}\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)\\ket{n_a}\\,.\n\\end{equation}\nSince we are working within the rotating-wave approximation [Eq.~\\eqref{Hdip}], the number of excitations is conserved at all times; and the operator $\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)$ counts the excitations in part of the system (which may have a photonic and an atomic component). If $m\\neq n$, either the number of excitations in that part, or the number of excitations in the rest, will be different between $\\ket{n_a}$ and $\\ket{m_a}$. Thus, $\\bra{m_a}\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)\\ket{n_a}=\\bra{n_a}\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)\\ket{n_a}\\delta_{m,n}$.\n\n\nThus, coherences across the Fock basis in the input state do not play any role. The rectifying device is fully characterised by studying its behavior on each Fock state $\\ket{n}$ separately, as we are going to do next.\n\n\n\\subsection{Photon-number pulses}\n\nLet us now study rectification for Fock states \\eqref{state:Fockn}. The corresponding flux for our chosen pulse \\eqref{xitau} can be computed from $[\\hat{a}_\\tau(0),(A^{\\dagger})^n] = n \\xi (\\tau) (A^{\\dagger})^{n-1}$, which implies $\\hat{a}_\\tau(0)\\ket{n} = \\sqrt{n} \\xi(\\tau) \\ket{n-1}$, and finally \\begin{eqnarray}\nF&=&\\expval{\\hat{a}_\\tau^\\dagger(0)\\hat{a}_\\tau(0)}{n} = \\frac{n\\Omega}{2}\\,.\n\\end{eqnarray}\n\nIn the Hamiltonian \\eqref{Hdip}, the operator $\\hat{b}_{\\omega}$ evolves as\n\\begin{equation}\\label{btomega}\n\\begin{split}\n \\hat{b}_{\\omega} (t) =& \\;\\hat{b}_{\\omega} (0) + g_{\\omega}^{(1)} \\int_{0}^{t} \\dd t' \\hat{\\sigma}_{-}^{(1)}(t') e^{i(\\omega-\\omega_1)t'} \\\\ \n & + g_{\\omega}^{(2)} \\int_{0}^{t} \\dd t' \\hat{\\sigma}_{-}^{(2)}(t') e^{i[(\\omega-\\omega_2)t' + \\omega d \/ v_g]}\\;.\n\\end{split}\n\\end{equation} As an example of what we have to compute, the number of reflected photons when the light comes in from the left is\n\\begin{equation}\\label{Nb:coh}\n\\begin{split}\n N_{\\text{ref}}^{(n)} (t) =& \\int_{0}^{\\infty}\\dd\\omega\\bra{n_a}\\hat{b}^{\\dagger}_{\\omega} (t)\\hat{b}_{\\omega} (t)\\ket{n_a}\\;,\\\\\n =& \\frac{\\gamma}{2} \\int_{0}^{t}\\dd t' \\Big\\{ \\expval{\\hat{\\sigma}_z^{(1)}(t')} + \\expval{\\hat{\\sigma}_z^{(2)}(t')} + 2 \\\\\n & + 2 \\left[ \\expval{\\hat{\\sigma}_{-}^{(1)}(t')\\hat{\\sigma}_{+}^{(2)}(t')} e^{-i (\\Delta_{12} t' - \\omega_2 \\mu)} + c.c.\\right] \n \\Big\\} \\;,\n\\end{split}\n\\end{equation}\nwhere $\\Delta_{12} =\\omega_1 - \\omega_2$ and $\\mu = d \/ v_g$.\n\nThe dynamics is given by the closed set of Heisenberg equations of motion presented in \\cite{Alex}, which we reproduce for completeness in Appendix \\ref{app:heis}. We solve them numerically in the steady state regime, for various input states, to obtain the desired expectation values \\eqref{ss:assume}.\n\nIn Fig.~\\ref{fig:eff}, the rectification factor for various input states is illustrated. In the case of a fixed bandwidth, $\\gamma\/\\Omega = 10^{2} $, we can see that we have a significant rectification for all the Fock states. For all these cases we observe a maximum rectification near 66\\%. However, the range of inter-atomic distances, $d$, and detunings, $\\Delta$, for which the strongest directionalities are observed, is narrower for Fock states with smaller number of photons. On the other hand, when we decrease the bandwidth, the directionality is lost for all the values of the distance between the qubits, $d$. For the case of a single photon, we observe that the rectification fades for a bandwidth between $\\gamma\/\\Omega \\sim 10^{3}$ and $10^{4}$. For an increasing number of photons, the rectification appears to be more robust against the narrowing of the bandwidth. For instance, in the case of 5 photons, the rectification of the device vanishes between $\\gamma\/\\Omega \\sim 10^{4}$ and $10^{5}$. Also, a similar behaviour can be observed for a fixed inter-atomic distance while varying the detuning. This is in agreement with the result in \\cite{Alex} that the device becomes symmetric for single photon inputs within the monochromatic limit.\n\nInterestingly, we observe that the rectification factor takes negative values (e.g. near $\\theta\/2\\pi = 0.5025$ and $\\Delta\/\\gamma = 0.04$). According to the definition \\eqref{DiodeEff}, this means that the direction of the diode became reversed. In Fig.~\\ref{fig:negeff}, we can see the maximum positive value of the rectification factor, for all the considered number states, are between 0.6 and 0.66. However, in the negative regime, the maximum efficiency of the device seems to depend on the number of photons. The strongest negativity is observed for the case of 22 photons. In other words, increasing or decreasing the number of photons results in weaker negative rectifications. Note that in this optimal case, $n=22$, the behaviour of the device is anti-symmetric. Hence, by changing the detuning between positive and negative values, one can flip the diode's preferred direction.\n\n\n\n\\section{Conclusions}\n\nIn this paper, we have clarified (and occasionally rectified, pun intended) the rectifying functionality of the simple optical diode sketched in Fig.~\\ref{fig:diode}. We have found that its behavior is completely determined by the behavior on the Fock states, and that some single-photon rectification does happen when the pulses are of finite length.\n\nThis study was done under two main assumptions: (i) that the pulses are monomode and long compared to decay time of atomic excitations, and (ii) that the rotating-wave approximation holds. To go beyond the first and consider composite and\/or short pulses, one would have to solve the time-dependent Heisenberg equations of motion. Removing the second assumption may be of interest too, given that the diode is naturally implemented with superconducting qubits \\cite{Clemens}, a platform in which ultrastrong coupling can be reached \\cite{Zueco,Niemczyk}.\n\n\n\\section*{Acknowledgment}\nThis research is supported by the National Research Foundation and the Ministry of Education, Singapore, under the Research Centres of Excellence programme. Part of this work was done when S.L.~was on exchange at the National University of Singapore in a NGNE programme. S.L.~acknowledges financial support from the School of Physics and the Qian Xuesen Honors College, Xi'an Jiaotong University.\n\n\n\\section{Introduction} \nControlling the photonic transport in integrated circuits is important for building future information and communication technologies. While a remarkable progress has been achieved on early implementations of such technologies \\cite{Asavanant,Larsen,Juan}, the development of an optical rectifier that can act on single or a few photon pulses remain challenging. These rectifying devices are useful for routing and isolating signals. Rectifiers based on ferromagnetic compounds have been demonstrated for increasing signal processing capabilities. However, they are inherently lossy and difficult to miniaturize and use in integrated circuits \\cite{Pozar}. Recently, using a semi-classical theory, a two atoms device has been identified as a candidate for exhibiting strong directionality \\cite{auffeves2014}. The transmission of light through this device would depend on the direction of propagation. If the input comes from one side it will pass through it, and if it comes from the other side it will get reflected. This has triggered a debate over whether such a device can perform with non-classical states of light \\cite{Alex,Auffeves2016,Vienna2016,Combes,Fang,Clemens,Gonzalez,Ordonez}.\n\nThis rectifying device, also referred to as \\textit{optical diode}, is consisting of a pair of two-level systems coupled to a one-dimensional waveguide. By changing the distance between the atoms and the detuning of their resonance frequencies, the efficiency of the diode was optimized. It has been shown that for light pulses in coherent states, the device can exhibit efficiencies of more than 60\\% \\cite{Alex,Auffeves2016,Vienna2016,Combes}. However, in the monochromatic limit, it has been found that this device cannot rectify single photon pulses \\cite{Alex}. Since significant directionality can be achieved for coherent inputs while a symmetric behaviour is predicted for single photon pulses, one might naturally ask: which components of the coherent state triggers the performance of the device? Do superpositions matter? How can we change these critical components, so we can improve the efficiency of this device for inputs with a few photon pulses? To answer this questions, we will use the same approach as in \\cite{Alex}. Using the Heisenberg equations of motion, we will show that by changing the length of the light pulse, we can reduce the number of photons necessary to activate the rectifying behaviour of the optical diode. \n\nThis article is organized as follow: in section II, we will introduce the theoretical model for this device and present how the rectification factor of the device is computed. In section III, we will discuss the cases of photon-number pulses. We will show that the main terms that will contribute to the rectification depend directly on the pulse duration. In section IV, we will present a few concluding remarks and discuss potential future directions of this research.\n\n\\section{The model}\n\n\\subsection{The rectifying device}\n\nAs illustrated in Fig.~\\ref{fig:diode}, the system of interest here is composed of a pair of atoms strongly coupled to a one dimensional waveguide. Each atom is model as a two-level system, for which $\\ket{g_j}$ and $\\ket{e_j}$ are the ground and excited states of the $j^{th}$ atom, respectively. The atom on the right is resonant with the central frequency of the light pulse $\\omega_0$. On the other hand, the atom on the left is detuned. We denote its resonance frequency by $\\omega_1$. Both frequencies are assumed to be much larger than the cutoff frequency of the waveguide \\cite{Snyder}. In this case, the dispersion relations read $\\omega = v_g \\abs{k}$, such that $k$ is the longitudinal wave-number of the field mode and $v_g$ is the group velocity of the light pulse in the waveguide. The free Hamiltonian of this system reads\n\\begin{equation} \\label{H0}\n \\hat{\\mathcal{H}}_{0} = \\sum_{j=1}^{2} \\hbar \\omega_j \\ketbra{e_j} + \\int_{0}^{\\infty} \\dd \\omega \\hbar \\omega (\\hat{a}^{\\dagger}_{\\omega}\\hat{a}^{}_{\\omega} + \\hat{b}^{\\dagger}_{\\omega}\\hat{b}^{}_{\\omega}) \\;,\n\\end{equation}\nsuch that the first term corresponds to the two atoms. The second term describes the field modes. The operator $\\hat{a}^{\\dagger}_{\\omega}$ ($\\hat{b}^{\\dagger}_{\\omega}$) creates a photon with frequency $\\omega$ moving towards the right (left). The interaction between the atoms and the propagating light pulse is described by the dipole Hamiltonian within the rotating wave approximation \\cite{Snyder,Roulet}\n\\begin{equation} \\label{Hdip}\n\\begin{split}\n \\hat{\\mathcal{H}}_{\\mathrm{dip}} =& -i \\hbar \\sum_{j=1}^{2} \\int_{0}^{\\infty} \\dd \\omega g_{\\omega}^{(j)} \\\\\n \\times \\Big[ & \\hat{\\sigma}_{+}^{(j)} \\left( \\hat{a}_{\\omega} e^{i \\omega \\frac{ x_j}{v_g}} + \\hat{b}_{\\omega} e^{- i \\omega \\frac{ x_j}{v_g}}\\right)e^{ - i (\\omega - \\omega_j)t} - \\textbf{H.c.} \\Big] \\;.\n\\end{split}\n\\end{equation}\nHere $x_j$, $g_{\\omega}^{(j)}$ and $\\hat{\\sigma}_{+}^{(j)} = \\ketbra{e_j}{g_j} $ are the position, the coupling strength and raising operator of the $j^{th}$ atom, respectively. Similarly to the previous study \\cite{Alex}, we assume the Weisskopf-Wigner approximation \\cite{Scully}, i.e. both atoms have the same coupling to the waveguide, $g_{\\omega}^{(1)} = g_{\\omega}^{(2)} = g$. Thus, the decay rate to the waveguide for both atoms is $\\gamma = 2\\pi g^2$.\n\n\\subsection{Computing rectification}\n\nSeveral figures of merit have been used in previous papers to capture rectification: we review them in Appendix \\ref{app:defs}. Here we define the rectifying factor of the device as\n\\begin{equation}\\label{DiodeEff}\n \\mathcal{R} = T_{\\rightarrow} - T_{\\leftarrow}\\,,\n\\end{equation} where $T_{\\rightarrow}$ is the transmittivity of the light coming from the left (i.e.~the fraction of left-incoming light scattered towards the right); analogously, $T_{\\leftarrow}$ is the fraction of right-incoming light scattered towards the left. Clearly $-1\\leq\\mathcal{R}\\leq 1$, the sign determining the direction in which the diode rectifies. Explicitly,\n\\begin{eqnarray}\\label{eqsT}\nT_{\\rightarrow}&=&\\frac{1}{F}\\lim_{t\\rightarrow\\infty}N_a^{\\rightarrow}(t)\\,=\\,\\frac{1}{F}(1-\\lim_{t\\rightarrow\\infty}N_b^{\\rightarrow}(t))\\,,\\\\\nT_{\\leftarrow}&=&\\frac{1}{F}\\lim_{t\\rightarrow\\infty}N_b^{\\leftarrow}(t)\n\\end{eqnarray}\nwhere $F$ is the mean photon flux per unit time, and where \\begin{eqnarray*}\nN_{b}^{\\textrm{input}} (t) = \\int_{0}^{\\infty} \\dd \\omega \\textrm{Tr}\\left[\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}^{}_{\\omega}(t)\\,\\rho_{\\textrm{input}}\\right]\\;,\n\\end{eqnarray*} (and similar definitions for similar quantities).\n\nIn this paper, we shall consider monomode Fourier-limited pulses. The mode that defines a pulse coming from the left is \n\\begin{equation}\\label{hatA}\n\\hat{A}^\\dagger = \\int_{0}^{\\infty} \\dd\\omega f (\\omega) \\hat{a}_{\\omega}^{\\dagger} = \\int_{0}^{\\infty} \\dd\\tau \\xi (\\tau) \\hat{a}_{\\tau}^{\\dagger} \\;,\n\\end{equation}\nwhere\n\\begin{equation}\n \\begin{split}\n \\hat{a}_{\\tau}^{} =& \\frac{1}{\\sqrt{2\\pi}} \\int \\dd\\omega \\hat{a}_{\\omega}^{} e^{- i (\\omega - \\omega_0) \\tau} \\;,\\\\\n \\xi (\\tau) =& \\frac{1}{\\sqrt{2\\pi}} \\int \\dd\\omega f (\\omega) e^{- i (\\omega - \\omega_0) \\tau}\\,.\n \\end{split}\n\\end{equation} For pulses coming from the right, we use the identical definition of a mode $\\hat{B}$, replacing the operators $(\\hat{a}_{.},\\hat{a}^\\dagger_{.})$ with $(\\hat{b}_{.},\\hat{b}^\\dagger_{.})$.\n\nLike in previous works, for definiteness the calculations will assume a square pulse\n\\begin{equation}\\label{xitau}\n\\xi (\\tau) = \\Bigg\\{\n \\begin{array}{cl}\n \\sqrt{\\Omega\/2} & \\text{for } 0 \\leq \\tau \\leq 2\/\\Omega, \\\\\n 0 & \\text{otherwise}.\n \\end{array}\n\\end{equation}\nFor an input state $\\ket{\\psi}$, the mean flux of photons coming from the left is \n\\begin{equation}\\label{flux}\n F = \\textrm{Tr}\\left[\\hat{a}_\\tau^\\dagger(0)\\hat{a}_\\tau(0)\\rho_{\\textrm{input}}\\right]\\;,\n\\end{equation}\n\nA basis of the total states of the system is denoted $\\ket{n_a,n_b,s_1,s_2}$ where $n_{a,b}$ is a Fock state in mode $\\hat{A}$ or $\\hat{B}$, while $s_j\\in\\{g,e\\}$ is the state of the $j$-th atom. Initially, the pulse will come from either the left or the right; and both atoms are assumed to be in the ground state to ensure that the rectifying device is passive. Thus our input states will always be of the form $\\ket{\\psi_a}\\equiv\\ket{\\psi,0,g,g}$, or $\\ket{\\psi_b}\\equiv\\ket{0,\\psi,g,g}$.\n\nInstead of solving the time-dependent equations of motion and sending $t\\rightarrow\\infty$, we shall consider only long pulses and thus assume that the system is in the steady state:\n\\begin{eqnarray}\n\\lim_{t\\rightarrow\\infty}N_b^{\\textrm{input}}(t)&\\approx&\\int_{0}^{\\infty} \\dd \\omega \\expval{\\hat{b}^{\\dagger}_{\\omega}\\hat{b}^{}_{\\omega}}_{ss,\\textrm{input}}\\,.\\label{ss:assume}\n\\end{eqnarray}\n\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=0.5\\textwidth,trim={0 9cm 0 0cm},clip]{figures\/Diode.pdf}\n \\caption{A scheme of the rectifying device consisting of two 2-level systems coupled to a one-dimensional waveguide. The one on the right is resonant with the central frequency of the light pulse, $\\omega_0$. The atom on the left is detuned, $\\omega_1 \\neq \\omega_0$.}\n \\label{fig:diode}\n\\end{figure}\n\n\\subsection{The case for pulses of finite length}\n\nAll previous fully-quantum studies of this rectifying device \\cite{Alex,Vienna2016,Auffeves2016} concluded that there is no rectification for single-photon Fock states. This conclusion was drawn in the limit of pulses of \\textit{infinite length}. One of the results of the next Section will be that \\textit{the device does show some rectification for single-photon Fock states} for pulses of finite length.\n\nThose same works also showed that high rectification is expected for coherent states, in qualitative (if not exact quantitative) agreement with a previous semi-classical study \\cite{auffeves2014}. Specifically, rectification was predicted \\cite{Alex} and later observed \\cite{Clemens} to have a high-value plateau when $F\/\\gamma$ ranges between $10^{-3}$ and $10^{-1}$. Let us now describe this case in more detail, to understand what may be the origin of the rectification.\n\nEvery coherent state is ultimately monomode \\cite{Alex90s,Alex90ss}: a coherent state pulse with average number of photons $\\bar{n}$, propagating towards the right, can be written as\n\\begin{equation}\\label{state:cohbis}\n \\ket{\\alpha_a} = \\exp\\big(\\sqrt{\\bar{n}} (\\hat{A}^\\dagger - \\hat{A})\\big) \\ket{\\text{\\o{}}}\\;.\n\\end{equation} Since $\\hat{a}_\\tau(0)\\ket{\\alpha} = \\sqrt{\\bar{n}}\\, \\xi(\\tau) \\ket{\\alpha}$, the flux \\eqref{flux} is given by $F=\\frac{\\bar{n} \\Omega}{2}$. If the device performance is maximized for plateau centred around $F\/\\gamma \\approx 10^{-2}$, then the optimal average number of photons in the pulse is\n\\begin{equation}\n \\bar{n}_{\\textrm{opt}} \\approx \\frac{2 \\gamma }{\\Omega} \\times 10^{-2}\\,.\n\\end{equation}\nIf we were to consider infinitely long pulses, as used in the single-photon case, then we would have to take the monochromatic limit $\\Omega \\rightarrow 0$; in which case $\\bar{n}_{\\textrm{opt}}$ diverges.\n\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=0.23\\textwidth]{figures\/f1s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p1pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f2s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p2pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f3s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p3pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f4s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p4pow.pdf}\\\\\n \\includegraphics[width=0.23\\textwidth]{figures\/f5s2.pdf}\n \\includegraphics[width=0.23\\textwidth]{figures\/p5pow.pdf}\\\\\n \n \\includegraphics[width=0.23\\textwidth]{figures\/barre.pdf}\n \\caption{The diode rectification $\\mathcal{R}$ for different Fock states ($n=1,2,3,4$ and 5 photons, respectively from top to bottom). Left panels: the rectification factors are computed as a function of the detuning $\\Delta = \\omega_0 - \\omega_1$, and the phase $\\theta = \\omega_0 d\/v_g$ while the bandwidth of the input pulse is fixed, ($\\gamma\/\\Omega = 10^{2}$. On the right panels, we plot the rectification factors for a fixed detuning, $\\Delta\/\\gamma = 0.1$ while we vary the bandwidth $\\Omega$.}\n \\label{fig:eff}\n \\vspace{-2cm}\n\\end{figure}\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=0.48\\textwidth]{figures\/RectDelta.pdf}\n \\vspace{-2mm}\n \\caption{The diode rectification $\\mathcal{R}$ for different Fock states ($n=1,2,3,4, 5$ and 22 photons). The pulse length is fixed ($\\Omega\/\\gamma = 10^{-2}$). The rectification factors are computed as a function of the detuning $\\Delta = \\omega_0 - \\omega_1$. For the case of $n=1$, the phase $\\theta$ is optimized in order to have the strongest negativity of the rectification factor. Ergo, we set $\\theta \/ 2\\pi = 0.5025$. Then we used the same value for the other cases. Here, we show the case of $n=22$ because for this Fock state we observed the strongest negative rectification.}\n \\label{fig:negeff}\n\\end{figure}\n\n\\section{Beyond the monochromatic case}\n\nIn this section, we consider the case of a pulse with a finite length. However, for our approximation \\eqref{ss:assume} to be valid, we need the pulse to be long enough such that the transient regime remains negligible. We shall set the bandwidth to be two orders of magnitude smaller than the decay rate to the waveguide, $\\Omega\/\\gamma = 10^{-2}$. For such a value, the necessary average number of photons to be in the middle of the plateau for the coherent state should be $\\bar{n}_{\\textrm{opt}} \\approx 2$.\n\n\\subsection{The device is not sensitive to coherences between Fock states}\n\nWe first prove that coherences between Fock states don't play any role in the dynamics of the rectification device. Consider as input state a generic superposition\n\\begin{eqnarray}\n\\ket{\\psi}&=&\\sum_{n=0}^\\infty c_n\\ket{n} \\;,\n\\end{eqnarray} \nwith $\\ket{n}$ standing for the Fock state with $n$ photons. For definiteness, we look at a special case, but the argument is the same for any of the calculations of interest here.\n\nWe assume therefore that the state $\\ket{\\psi}$ is prepared in mode $\\hat{A}$, i.e.~the input state is $\\ket{\\psi_a}$ with the Fock states given by \\cite{Alex90s,Alex90ss}\n\\begin{equation}\\label{state:Fockn}\n \\ket{n} = \\frac{(\\hat{A}^{\\dagger})^n}{\\sqrt{n!}} \\ket{\\text{\\o{}}}\\,.\n\\end{equation} Then we look at the number of reflected photons\n\\begin{equation}\\label{eq:NnmAll}\n N_{b}^{\\psi} (t) = \n \\sum_{m,n}c_nc_m^*\\int_{0}^{\\infty}\\dd\\omega\\bra{m_a}\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)\\ket{n_a}\\,.\n\\end{equation}\nSince we are working within the rotating-wave approximation [Eq.~\\eqref{Hdip}], the number of excitations is conserved at all times; and the operator $\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)$ counts the excitations in part of the system (which may have a photonic and an atomic component). If $m\\neq n$, either the number of excitations in that part, or the number of excitations in the rest, will be different between $\\ket{n_a}$ and $\\ket{m_a}$. Thus, $\\bra{m_a}\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)\\ket{n_a}=\\bra{n_a}\\hat{b}^{\\dagger}_{\\omega}(t)\\hat{b}_{\\omega}(t)\\ket{n_a}\\delta_{m,n}$.\n\n\nThus, coherences across the Fock basis in the input state do not play any role. The rectifying device is fully characterised by studying its behavior on each Fock state $\\ket{n}$ separately, as we are going to do next.\n\n\n\\subsection{Photon-number pulses}\n\nLet us now study rectification for Fock states \\eqref{state:Fockn}. The corresponding flux for our chosen pulse \\eqref{xitau} can be computed from $[\\hat{a}_\\tau(0),(A^{\\dagger})^n] = n \\xi (\\tau) (A^{\\dagger})^{n-1}$, which implies $\\hat{a}_\\tau(0)\\ket{n} = \\sqrt{n} \\xi(\\tau) \\ket{n-1}$, and finally \\begin{eqnarray}\nF&=&\\expval{\\hat{a}_\\tau^\\dagger(0)\\hat{a}_\\tau(0)}{n} = \\frac{n\\Omega}{2}\\,.\n\\end{eqnarray}\n\nIn the Hamiltonian \\eqref{Hdip}, the operator $\\hat{b}_{\\omega}$ evolves as\n\\begin{equation}\\label{btomega}\n\\begin{split}\n \\hat{b}_{\\omega} (t) =& \\;\\hat{b}_{\\omega} (0) + g_{\\omega}^{(1)} \\int_{0}^{t} \\dd t' \\hat{\\sigma}_{-}^{(1)}(t') e^{i(\\omega-\\omega_1)t'} \\\\ \n & + g_{\\omega}^{(2)} \\int_{0}^{t} \\dd t' \\hat{\\sigma}_{-}^{(2)}(t') e^{i[(\\omega-\\omega_2)t' + \\omega d \/ v_g]}\\;.\n\\end{split}\n\\end{equation} As an example of what we have to compute, the number of reflected photons when the light comes in from the left is\n\\begin{equation}\\label{Nb:coh}\n\\begin{split}\n N_{\\text{ref}}^{(n)} (t) =& \\int_{0}^{\\infty}\\dd\\omega\\bra{n_a}\\hat{b}^{\\dagger}_{\\omega} (t)\\hat{b}_{\\omega} (t)\\ket{n_a}\\;,\\\\\n =& \\frac{\\gamma}{2} \\int_{0}^{t}\\dd t' \\Big\\{ \\expval{\\hat{\\sigma}_z^{(1)}(t')} + \\expval{\\hat{\\sigma}_z^{(2)}(t')} + 2 \\\\\n & + 2 \\left[ \\expval{\\hat{\\sigma}_{-}^{(1)}(t')\\hat{\\sigma}_{+}^{(2)}(t')} e^{-i (\\Delta_{12} t' - \\omega_2 \\mu)} + c.c.\\right] \n \\Big\\} \\;,\n\\end{split}\n\\end{equation}\nwhere $\\Delta_{12} =\\omega_1 - \\omega_2$ and $\\mu = d \/ v_g$.\n\nThe dynamics is given by the closed set of Heisenberg equations of motion presented in \\cite{Alex}, which we reproduce for completeness in Appendix \\ref{app:heis}. We solve them numerically in the steady state regime, for various input states, to obtain the desired expectation values \\eqref{ss:assume}.\n\nIn Fig.~\\ref{fig:eff}, the rectification factor for various input states is illustrated. In the case of a fixed bandwidth, $\\gamma\/\\Omega = 10^{2} $, we can see that we have a significant rectification for all the Fock states. For all these cases we observe a maximum rectification near 66\\%. However, the range of inter-atomic distances, $d$, and detunings, $\\Delta$, for which the strongest directionalities are observed, is narrower for Fock states with smaller number of photons. On the other hand, when we decrease the bandwidth, the directionality is lost for all the values of the distance between the qubits, $d$. For the case of a single photon, we observe that the rectification fades for a bandwidth between $\\gamma\/\\Omega \\sim 10^{3}$ and $10^{4}$. For an increasing number of photons, the rectification appears to be more robust against the narrowing of the bandwidth. For instance, in the case of 5 photons, the rectification of the device vanishes between $\\gamma\/\\Omega \\sim 10^{4}$ and $10^{5}$. Also, a similar behaviour can be observed for a fixed inter-atomic distance while varying the detuning. This is in agreement with the result in \\cite{Alex} that the device becomes symmetric for single photon inputs within the monochromatic limit.\n\nInterestingly, we observe that the rectification factor takes negative values (e.g. near $\\theta\/2\\pi = 0.5025$ and $\\Delta\/\\gamma = 0.04$). According to the definition \\eqref{DiodeEff}, this means that the direction of the diode became reversed. In Fig.~\\ref{fig:negeff}, we can see the maximum positive value of the rectification factor, for all the considered number states, are between 0.6 and 0.66. However, in the negative regime, the maximum efficiency of the device seems to depend on the number of photons. The strongest negativity is observed for the case of 22 photons. In other words, increasing or decreasing the number of photons results in weaker negative rectifications. Note that in this optimal case, $n=22$, the behaviour of the device is anti-symmetric. Hence, by changing the detuning between positive and negative values, one can flip the diode's preferred direction.\n\n\n\n\\section{Conclusions}\n\nIn this paper, we have clarified (and occasionally rectified, pun intended) the rectifying functionality of the simple optical diode sketched in Fig.~\\ref{fig:diode}. We have found that its behavior is completely determined by the behavior on the Fock states, and that some single-photon rectification does happen when the pulses are of finite length.\n\nThis study was done under two main assumptions: (i) that the pulses are monomode and long compared to decay time of atomic excitations, and (ii) that the rotating-wave approximation holds. To go beyond the first and consider composite and\/or short pulses, one would have to solve the time-dependent Heisenberg equations of motion. Removing the second assumption may be of interest too, given that the diode is naturally implemented with superconducting qubits \\cite{Clemens}, a platform in which ultrastrong coupling can be reached \\cite{Zueco,Niemczyk}.\n\n\n\\section*{Acknowledgment}\nThis research is supported by the National Research Foundation and the Ministry of Education, Singapore, under the Research Centres of Excellence programme. Part of this work was done when S.L.~was on exchange at the National University of Singapore in a NGNE programme. S.L.~acknowledges financial support from the School of Physics and the Qian Xuesen Honors College, Xi'an Jiaotong University.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\t\n\nErd\\H{o}s \\cite{Erdos} in 1946 asked the question of finding the minimal number of distinct distances among any $N$ points in the plane. The breakthrough work of Guth-Katz \\cite{Guth-Katz} gave the lower bound $\\geq C\\frac{N}{\\log N}$ for some constant $C>0$ in the Euclidean plane, which is sharp up to a factor of $\\log$. Another related and widely studied conjecture is\nthe Falconer's conjecture which asks about the lower bound of the Hausdorff dimension of the sets in $\\mathbb{R}^d$ for which the difference set has positive Lebesgue measure. The Falconer's conjecture can be viewed as a continuous analogue of the distinct distances problem. Interested readers may check Falconer \\cite{Falconer}, Guth-Iosevich-Ou-Wang \\cite{Guth-Iosevich-Ou-Wang}, Iosevich \\cite{Iosevich} etc. \nThe Erd\\H{o}s-Falconer type problems have been generalized to other spaces and applied to certain sum-product estimates, see e.g. Bourgain-Tao \\cite{Bourgain-Tao}, Hart-Iosevich-Koh-Rudnev \\cite{HIKR}, Roche-Newton and Rudnev \\cite{RocheNewton-Rudnev}, Rudnev-Selig \\cite{Rudnev-Selig}, Sheffer-Zahl \\cite{Sheffer-Zahl}, and blog of Tao \\cite{Tao} etc.\n However, the distinct distances problem has not been considered in hyperbolic surfaces until very recently by Lu and the author in \\cite{Lu-Meng} where the modular surface and hyperbolic surfaces with cocompact fundamental groups are studied. But this problem is still open for more general hyperbolic surfaces arising from non-cocompact Fuchsian groups. \n\nIn this paper, for all cofinite Fuchsian groups $\\Gamma$, we give complete answer to the distinct distances problem for all hyperbolic surfaces $\\Gamma\\backslash\\mathbb{H}^2$ endowed with the hyperbolic metric from $\\mathbb{H}^2$. \n\n\n\\begin{theorem}\\label{thm-cofinite}\n\tFor any cofinite Fuchsian group $\\Gamma\\subset \\mathrm{PSL}(2, \\mathbb{R})$, any set of $N$ points on the hyperbolic surface $\\Gamma\\backslash\\mathbb{H}^2$ determines $\\geq C_{\\Gamma}\\frac{N}{\\log N}$ distinct distances for some constant $C_{\\Gamma}$ depending only on $\\Gamma$. \n\\end{theorem}\n\n\nIn particular, for finite index subgroups of the modular group $\\mathrm{PSL}(2, \\mathbb{Z})$, we extract out the dependence of the implied constants on the index. \n\n\n\n\n\\begin{theorem}\\label{thm-finite-index}\n\tFor any finite index subgroup $\\Gamma$ of $\\mathrm{PSL}(2, \\mathbb{Z})$ with $[ \\mathrm{PSL}(2,\\mathbb{Z}):\\Gamma ]=\\mu$, any set of $N$ points on the hyperbolic surface $\\Gamma\\backslash\\mathbb{H}^2$ determines $\\geq C \\frac{N}{\\mu\\log N}$ distinct distances for some absolute constant $C>0$. \n\\end{theorem}\n\nTheorem \\ref{thm-finite-index} has application to equilateral dimension problem. The equilateral dimension of a metric space is the maximal number of points in the space with pairwise equal distance. It has been studied in various spaces, see Alon-Milman \\cite{Alon-Milman}, Guy \\cite{Guy}, Koolen \\cite{Koolen} etc. For instance, the equilateral dimension of the $n$-dimensional Euclidean space is $n+1$. However, we are not aware of any result in literature about the equilateral dimension of general hyperbolic surfaces. We observe that the lower bound in Theorem \\ref{thm-finite-index} is not trivial for distinct distances among any set of size $N\\gg \\mu^{1+\\epsilon}$. Thus the following corollary holds.\n\\begin{corollary}\n\tFor any subgroup $\\Gamma$ of $\\mathrm{PSL}(2, \\mathbb{Z})$ with finite index $[ \\mathrm{PSL}(2,\\mathbb{Z}):\\Gamma ]=\\mu$, the equilateral dimension of the hyperbolic surface $\\Gamma\\backslash\\mathbb{H}^2$ is $\\ll \\mu^{1+\\epsilon}$ for any $ \\epsilon>0$. \n\\end{corollary}\n\nThe isometry group of the hyperbolic plane $\\mathbb{H}^2$ is $\\mathrm{PSL}(2, \\mathbb{R})$ which acts on $\\mathbb{H}^2$ by M\\\"{o}bius transformation:\n\\[z\\mapsto \\gamma(z):=\\frac{az+b}{cz+d}, \\text{ for }\\gamma=\\begin{pmatrix}a&b\\\\c&d\\end{pmatrix}\\in\\mathrm{PSL}(2, \\mathbb{R}), z\\in\\mathbb{H}^2.\\] \nFor any discrete subgroup $\\Gamma$ of $\\mathrm{PSL}(2, \\mathbb{R})$, i.e. a \\textit{Fuchsian group}, the distance between any two points $p, q $ on the hyperbolic surface $ Y\\cong \\Gamma\\backslash\\mathbb{H}^2$ is \n$$d_Y(p, q):=\\min_{\\gamma_1, \\gamma_2\\in \\Gamma} d_{\\mathbb{H}^2}(\\gamma_1(p), \\gamma_2(q))=\\min_{\\gamma_1, \\gamma_2\\in\\Gamma} d_{\\mathbb{H}^2}(p, \\gamma_1^{-1}\\gamma_2(q) )=\\min_{\\gamma\\in\\Gamma}d_{\\mathbb{H}^2}(p, \\gamma(q)).$$\n\\begin{figure}[ht]\n\t\\centering\n\t\\includegraphics[width=0.6\\textwidth]{surface-show.png}\n\t\\caption{Distances on hyperbolic surface}\n\\end{figure}\n\nInstead of calculating distances on the surface directly, we consider representatives of the points in a fundamental domain $F_{\\Gamma}$ of $\\Gamma$. In \\cite{Lu-Meng}, Lu and the author introduced the concept of a ``geodesic cover\" $\\Gamma'\\subset\\Gamma$ such that for any $p, q\\in F_{\\Gamma}$, \n$$ d_Y(p, q)= d_{\\mathbb{H}^2}(p, \\gamma' q) ~\\text{for some}~\\gamma'\\in\\Gamma'.$$\nIf there exists a finite geodesic cover, one can derive a lower bound for the distinct distance problem on the hyperbolic surface $\\Gamma\\backslash\\mathbb{H}^2$. For the modular surface $\\mathrm{PSL}(2, \\mathbb{Z})\\backslash\\mathbb{H}^2$, we would be able to find a finite geodesic cover by working explicitly with matrices in $\\mathrm{PSL}(2, \\mathbb{Z})$. However, it is hard to tackle general non-cocompact Fuchsian groups this way since we cannot explicitly write out all the elements. Another difficulty to find such a finite geodesic cover is, the number of representatives we need to examine would blow up if the fundamental domain has many inequivalent cusps. This is not an issue for modular surface which has only one inequivalent cusp, and that the imaginary parts of points in a fundamental domain of $\\mathrm{PSL}(2, \\mathbb{Z})$ are all bounded below (or bounded above if we choose other type of fundamental domain). Therefore, representatives we have to examine will not have very small imaginary parts and the number of them could be bounded. But in the general case, if a pair of points are close to two inequivalent cusps respectively, the number of representatives we have to examine might lose control. \n\n\nIn order to overcome such difficulties, we propose a more general concept of a geodesic cover defined on any subset of a fundamental domain $F_{\\Gamma}$ and also defined in different base groups, see Definition \\ref{defn-geodesic-covering-number}. By building relations between geodesic covers of different subregions in $F_{\\Gamma}$ and geodesic covers of certain regions in different groups, we prove lower bounds for distinct distances on hyperbolic surfaces associated with any cofinite Fuchsian group. See Lemmas \\ref{lem-geodesic-number-Distinct-distance}, \\ref{lem-geodesic-cover-number-finite} and \\ref{lem-finite-index-geo-cover-number} for details. \n\n\n\n\n\n\n\\bigskip\n\\textbf{Acknowledgment.} The author is partially supported by the Humboldt Professorship of Professor Harald Helfgott. \n\n\n\\section{Preliminaries and preparations}\n\nFirst we briefly summarize the properties of Fuchsian groups (see Beardon \\cite{Beardon} or Katok \\cite{Katok} for more related materials).\nA subgroup $\\Gamma$ of $\\mathrm{PSL}(2, \\mathbb{R})$ is a \\textit{Fuchsian group} if and only if $\\Gamma$ acts \\textit{properly discontinuously} on $\\mathbb{H}^2$. Thus the $\\Gamma$-orbit of any point $z\\in\\mathbb{H}^2$ is\\textit{ locally finite}, which means any compact set $K\\subset\\mathbb{H}^2$ contains only finite number of orbit points, i.e. the set $\\Gamma z\\cap K$ is finite for any $z\\in\\mathbb{H}^2$.\n\nA \\textit{cofinite} Fuchsian group is a discrete subgroup of $\\mathrm{PSL}(2, \\mathbb{R})$ of finite covolume i.e. a fundamental domain of $\\Gamma\\backslash\\mathbb{H}^2$ has finite hyperbolic area. A cofinite discrete subgroup is also called a \\textit{lattice} in some other contexts. \nSiegel's theorem (see \\cite{Katok}, Theorem 4.1.1) says cofinite Fuchsian group is \\textit{geometrically finite}, i.e. there exists a convex fundamental domain with finitely many sides.\n\n\n\n\n\n\n\n\n\nThe cocompact Fuchsian groups has been considered in \\cite{Lu-Meng}. In this paper, we focus on non-cocompact case. Suppose $\\Gamma$ has parabolic elements, and thus its fundamental domain $F_{\\Gamma}$ must have a vertex on $\\hat{\\mathbb{R}}$ which is called a \\textit{cusp}. Since we assume $\\Gamma$ is cofinite, by Siegel's theorem, its fundamental domain $F_{\\Gamma}$ has finitely many cusps.\n\n\nWe use an idea of Iwaniec (see \\cite{Iwaniec}, \\S 2.2) to partition the fundamental domain of a Fuchsian group. Define the stability group as\n$$\\Gamma_z:=\\{ \\gamma\\in\\Gamma: \\gamma z=z \\}.$$\nGiven a cusp $\\mathfrak{a}\\in \\hat{\\mathbb{R}}$ for $\\Gamma$. The stability group $\\Gamma_{\\mathfrak{a}}$ is a cyclic group generated by a parabolic element, say $\\Gamma_{\\mathfrak{a}}=\\langle \\gamma_{\\mathfrak{a}}\\rangle$. There exists $\\sigma_{\\mathfrak{a}}\\in SL(2, \\mathbb{R})$ such that\n\\begin{equation}\\label{eq-conjugate-to-translation}\n\\sigma_{\\mathfrak{a}}\\infty=\\mathfrak{a}, \\quad \\sigma_{\\mathfrak{a}}^{-1}\\gamma_{\\mathfrak{a}}\\sigma_{\\mathfrak{a}}=\\begin{pmatrix}1&1\\\\ 0&1\\end{pmatrix}.\n\\end{equation}\nThen $\\sigma_{\\mathfrak{a}}^{-1}$ sends $\\mathfrak{a}$ to $\\infty$ and $\\sigma_{\\mathfrak{a}}$ maps the strip\n\\begin{equation}\\label{infinite-part-of-fundamental-domain}\nP(T):=\\{ z=x+iy: 00$ depending on $\\Gamma$. \t\n\t\n\\textbf{Case 2).} There are more than $N\/2$ points on $F_{\\bowtie,T}$. Let $n_c<\\infty$ be the number of cusps for the fundamental domain $F_{\\Gamma}$. Then there exists one cusp $\\mathfrak{b}$ such that $F_{\\mathfrak{b},T}$ contains more than $N\/2n_c$ points. We may assume all these points lie in the interior of $F_{\\mathfrak{b}, T}$. Denote the set of points on $F_{\\mathfrak{b},T}$ by $\\mathcal{S}_2$. By a similar argument as in Case 1), we deduce that\n\t\\begin{equation}\n\t|d_Y(\\mathcal{S})|\\geq |d_Y(\\mathcal{S}_2)|\\gg \\frac{N\/n_c}{ K^3_{\\Gamma}(F_{\\mathfrak{b}, T}) (\\log(K_{\\Gamma}(F_{\\mathfrak{b},T}) ) +\\log N ) } \\geq C''_{\\Gamma}\\frac{N}{\\log N},\n\t\\end{equation}\n\tfor some constant $C''_{\\Gamma}>0$ depending on $\\Gamma$.\n\t\n\tCombining the cases 1) and 2), we finish the proof.\n\\end{proof}\n\n\n\n\n\n\n\n\\section{ Geodesic-covering numbers for cofinite Fuchsian groups }\n\n\nIn this section, we give the proof of Theorem \\ref{thm-cofinite} based on Lemma \\ref{lem-geodesic-number-Distinct-distance}. We only need to bound the geodesic-covering numbers of $F_T$ and $F_{\\mathfrak{a}, T} $ for every cusp $ \\mathfrak{a}$. \n\\begin{lemma}\\label{lem-geodesic-cover-number-finite}\n\tAssume $\\Gamma$ is a cofinite Fuchsian group with a fundamental domain $F_{\\Gamma}$. If we partition $F_{\\Gamma}$ as in \\eqref{eq-partition-cuspidal-central} for some large enough $T$ depending on $\\Gamma$, the geodesic-covering numbers of $F_T$ and $F_{\\mathfrak{a}, T}$ for every cusp $\\mathfrak{a}$ in $\\Gamma$ are all finite. This is also true for the closure of $F_T$, i.e. $K_{\\Gamma}(\\overline{F}_T)<\\infty$. \n\\end{lemma}\n\n\n\\begin{proof}\nWe need to know the basic shape of a fundamental domain for any fuchsian group. A convenient choice for us is \\textit{Ford domain} which was first introduced by L. R. Ford \\cite{Ford}. It is known that Ford domain is a fundamental domain (see \\cite{Katok}, Theorem 3.3.5). \nThere are concrete methods to construct fundamental domains of Fuchsian groups, interested readers may check Voight \\cite{Voight} for an algorithmic method, and Kulkarni \\cite{Kul} for construction of special polygons (also a fundamental domain) for subgroups of modular group using Farey symbol. \n\n\nLet $F_{\\Gamma}$ be a fundamental domain of cofinite $\\Gamma$ with finite number of sides and finite number of cusps. We partition $F_{\\Gamma}$ as in \\eqref{eq-partition-cuspidal-central} for some $T$ we choose later, \n$$F_{\\Gamma}=F_T\\bigcup_{\\mathfrak{a} ~{\\rm cusp}}F_{\\mathfrak{a}, T}. $$\n\n\n1) First we show that, for some $T$, the geodesic-covering number of $F_{\\mathfrak{a}, T}$ in $\\Gamma$ is finite for every cusp $\\mathfrak{a}$. In order to do this, we make use of Ford domains. \n\nFor any cusp $\\mathfrak{a}$, by \\eqref{eq-conjugate-to-translation}, there exists $\\sigma_{\\mathfrak{a}}$ such that the stability group $\\Gamma_{\\mathfrak{a}}$ is generated by \n$$\\sigma_{\\mathfrak{a}}\\begin{pmatrix}\n1 & 1\\\\\n0 & 1\n\\end{pmatrix}\\sigma^{-1}_{\\mathfrak{a}},$$\nand the fundamental domain of $\\sigma^{-1}_{\\mathfrak{a}}\\Gamma_{\\mathfrak{a}}\\sigma_{\\mathfrak{a}}$ is \n\\begin{equation}\nP:=\\{ z\\in\\mathbb{H}^2: 0\\leq x< 1, y>0\\}.\n\\end{equation}\nDenote $\\widetilde{\\Gamma}^{\\mathfrak{a}}:=\\sigma^{-1}_{\\mathfrak{a}}\\Gamma\\sigma_{\\mathfrak{a}}$ and $$\\widetilde{\\Gamma}_{\\infty}^{\\mathfrak{a}}:=\\sigma^{-1}_{\\mathfrak{a}}\\Gamma_{\\mathfrak{a}}\\sigma_{\\mathfrak{a}}=\\left\\langle \\begin{pmatrix}\n1 & 1\\\\\n0 & 1\n\\end{pmatrix} \\right\\rangle.$$\n\n\nBy \\eqref{infinite-part-of-fundamental-domain} and \\eqref{eq-F-cusp-to-F-infty}, the geodesic-covering number of $F_{\\mathfrak{a}, T}$ in $\\Gamma$ is the same as the geodesic-covering number of $\\sigma^{-1}_{\\mathfrak{a}}(F_{\\mathfrak{a}, T} )=P(T)$ in $\\sigma^{-1}_{\\mathfrak{a}}\\Gamma\\sigma_{\\mathfrak{a}}$, i.e. \n\\begin{equation}\\label{eq-geo-cover-number-to-P}\nK_{\\Gamma}(F_{\\mathfrak{a}, T})=K_{\\widetilde{\\Gamma}^{\\mathfrak{a}}}(P(T)).\n\\end{equation}\n\n\nWe define a domain associated with cusp $\\mathfrak{a}$ as\n\\begin{align}\\label{eq-Ford-domain-cusp}\n\\mathcal{D}_{\\mathfrak{a}}:=& \\{ z\\in P: {\\rm Im}(\\gamma z) <{\\rm Im} (z), \\forall \\gamma\\in \\widetilde{\\Gamma}^{\\mathfrak{a}}\\setminus\\widetilde{\\Gamma}_{\\infty}^{\\mathfrak{a}} \\}\\nonumber\\\\\n=&\\{ z\\in P: |cz+d|>1, \\forall \\begin{pmatrix}\n* & *\\\\\nc & d\n\\end{pmatrix}\\in \\widetilde{\\Gamma}^{\\mathfrak{a}}\\setminus\\widetilde{\\Gamma}_{\\infty}^{\\mathfrak{a}} \\}\n\\end{align}\nwhich is a \\textit{Ford domain} and thus a fundamental domain of $\\widetilde{\\Gamma}^{\\mathfrak{a}}$. Note that $\\sigma_{\\mathfrak{a}}^{-1}(F_{\\Gamma})$ may not be the same as $\\mathcal{D}_{\\mathfrak{a}}$. \n\n\nWe want to choose large enough $T$ such that $P(T)\\subset \\mathcal{D}_{\\mathfrak{a}}$ for all cusp $\\mathfrak{a}$. Since $\\mathcal{D}_{\\mathfrak{a}}$ is a fundamental domain of $ \\widetilde{\\Gamma}^{\\mathfrak{a}} $, the boundary of $\\mathcal{D}_{\\mathfrak{a}}$ consists of finite number of pieces from\\textit{ isometric circles} of the form $|z+\\frac{d}{c}|=\\frac{1}{|c|}$ for some \n$$c\\neq 0, \\begin{pmatrix}\n* & *\\\\\nc & d\n\\end{pmatrix}\\in \\widetilde{\\Gamma}^{\\mathfrak{a}}.$$\nThus there is a largest radius among these isometric circles, say $\\frac{1}{c_{\\mathfrak{a}}}$, actually (see \\cite{Iwaniec}, \\S 2.6) \n\\begin{equation}\\label{eq-min-c}\nc_{\\mathfrak{a}}=\\min\\left\\{ c>0: \\begin{pmatrix}\n* & *\\\\\nc & *\n\\end{pmatrix}\\in\\widetilde{\\Gamma}^{\\mathfrak{a}}\\setminus\\widetilde{\\Gamma}_{\\infty}^{\\mathfrak{a}} \\right\\}.\n\\end{equation}\nFor the fundamental domain $F_{\\Gamma}$, there are only finite number of cusps, we choose any large enough\n$$T\\geq 100+10\\max_{\\mathfrak{a} ~{\\rm cusp}} \\tfrac{1}{c_{\\mathfrak{a}} },$$\nthen $P(T)=\\sigma^{-1}_{\\mathfrak{a}}(F_{\\mathfrak{a}, T} )\\subset \\mathcal{D}_{\\mathfrak{a}}$ for every cusp $\\mathfrak{a}$. \n\nFor the above choice of $T$, we are ready to estimate $K_{\\widetilde{\\Gamma}^{\\mathfrak{a}}}(P(T))$ for any $\\mathfrak{a}$. Consider the set\n\\begin{equation}\n\\mathcal{A}:=\\big\\{ \\gamma\\in\\widetilde{\\Gamma}^{\\mathfrak{a}}: d_{\\mathbb{H}^2}(z_1, \\gamma z_2)\\leq d_{\\mathbb{H}}(z_1, z_2), z_1, z_2\\in P(T), {\\rm Im}(z_1)\\geq {\\rm Im}(z_2) \\big\\},\n\\end{equation}\nwhich, by Definition \\ref{defn-geodesic-covering-number}, is a geodesic cover of $P(T)$ in $\\widetilde{\\Gamma}^{\\mathfrak{a}}$. \n\nFor any two points $z_1=x_1+iy_1$ and $z_2=x_2+iy_2$ in $P(T)$ with $y_1\\geq y_2$, the only possible isometries $\\gamma$ from\n$$ \\widetilde{\\Gamma}_{\\infty}^{\\mathfrak{a}}=\\left\\langle \\begin{pmatrix}\n1 & 1\\\\\n0 & 1\n\\end{pmatrix} \\right\\rangle,$$\nsuch that $d_{\\mathbb{H}^2}(z_1, \\gamma z_2)\\leq d_{\\mathbb{H}^2}(z_1, z_2)$ are \n\\begin{equation}\n\\mathcal{T}:=\\left\\{ \\begin{pmatrix}\n1 & -1\\\\\n0 & 1\n\\end{pmatrix}, \n\\begin{pmatrix}\n1 &0\\\\\n0 & 1\n\\end{pmatrix},\n\\begin{pmatrix}\n1 & 1\\\\\n0 & 1\n\\end{pmatrix} \\right\\}.\n\\end{equation}\n\nIf $\\gamma\\in \\widetilde{\\Gamma}^{\\mathfrak{a}}\\setminus\\widetilde{\\Gamma}_{\\infty}^{\\mathfrak{a}} $, by the construction of $\\mathcal{D}_{\\mathfrak{a}}$ and \\eqref{eq-min-c}, we have\n\\begin{equation}\n{\\rm Im}(\\gamma z_2)=\\frac{y_2}{ (cx_2+d)^2+c^2 y_2^2}\\leq \\frac{1}{c^2 y_2}\\leq \\frac{1}{c^2_{\\mathfrak{a}} y_2}.\n\\end{equation}\nSince $y_2\\geq T\\geq 100+\\frac{10}{c_{\\mathfrak{a}}}$, we deduce that\n\\begin{equation}\n{\\rm Im}(\\gamma z_2)\\leq \\frac{1}{100 c^2_{\\mathfrak{a}}+10c_{\\mathfrak{a}}}<\\frac{1}{10c_{\\mathfrak{a}}}.\n\\end{equation}\nDenote $\\gamma z_2=x_0+iy_0$ ($y_0<\\frac{1}{10c_{\\mathfrak{a}}}$), then by the hyperbolic distance formula, \n\\begin{equation}\\label{eq-hyperbolic-cosh-distance}\n2\\cosh(d_{\\mathbb{H}^2}(z_1, z_2))=\\frac{(x_1-x_2)^2+y_1^2+y_2^2}{y_1 y_2},\n\\end{equation}\nand $|x_1-x_2|\\leq 1$, $y_1\\geq y_2\\geq T\\geq 100+\\frac{10}{c_{\\mathfrak{a}}}$, \nwe derive\n\\begin{align}\n&2\\cosh(d_{\\mathbb{H}^2}(z_1, \\gamma z_2))\n-2\\cosh(d_{\\mathbb{H}^2}(z_1, z_2)) \\nonumber\\\\\n&=\\frac{(x_1-x_0)^2+y_1^2+y_0^2}{y_1 y_0 }- \\frac{(x_1-x_2)^2+y_1^2+y_2^2}{ y_1 y_2} \\nonumber\\\\\n&\\geq \\frac{y_1}{y_0}-\\frac{1}{y_1 y_2}-\\frac{y_1}{y_2}-\\frac{y_2}{y_1} \\nonumber\\\\\n&\\geq y_1\\Big(\\frac{1}{y_0}-\\frac{1}{y_2}\\Big)-\\frac{1}{100^2}-1\\nonumber\\\\\n&\\geq \\frac{10}{c_{\\mathfrak{a}}} \\Big( 10c_{\\mathfrak{a}}-\\frac{c_{\\mathfrak{a}}}{10} \\Big)-2=99-2>0.\n\\end{align}\n\nHence we have $\\mathcal{A}=\\mathcal{T}$. We derive that the geodesic-covering number of $P(T)$ in $\\widetilde{\\Gamma}^{\\mathfrak{a}}$ is $\\leq 3$. Since our choice of $T$ works for all cusps, and by \\eqref{eq-geo-cover-number-to-P}, we conclude that the geodesic-covering number of $F_{\\mathfrak{a}, T}$ in $\\Gamma$ is finite for all cusp $\\mathfrak{a}$, precisely $K_{\\Gamma}(F_{\\mathfrak{a}, T})\\leq 3$. \n\n\n2) Now we bound the geodesic-covering number of the central part $F_T$ in $\\Gamma$. \nDefine the diameter of $F_T$ as\n$$\\diam(F_T):=\\max_{p, q\\in F_T} d_{\\mathbb{H}^2}(p, q).$$\nSince $F_T$ is bounded, the diameter $\\diam(F_T)$ is finite. Pick any point $O$ inside $F_T$ which is not fixed by any element in $\\Gamma$ except identity, then the set\n$$\\mathcal{B}:=\\big\\{ \\gamma\\in\\Gamma: d_{\\mathbb{H}^2}(O, \\gamma(O))\\leq 3\\diam(F_T) \\big\\}$$\nis a geodesic cover of $F_T$ in $\\Gamma$. Indeed, for any $\\gamma\\not\\in \\mathcal{B}$ and any two points $p, q\\in F_T$, by triangle inequality, we get\n\\begin{align}\nd_{\\mathbb{H}^2}(p, \\gamma(q))&\\geq d_{\\mathbb{H}^2}(O, \\gamma(O))-d_{\\mathbb{H}^2}(p, O)-d_{\\mathbb{H}^2}(\\gamma(O), \\gamma(q))\\nonumber\\\\\n&\\geq 3\\diam(F_T)-\\diam(F_T)-\\diam(F_T)=\\diam(F_T)\\geq d_{\\mathbb{H}^2}(p, q).\n\\end{align}\n\nSince a Fuchsian group $\\Gamma$ acts properly discontinuously on $\\mathbb{H}^2$, the $\\Gamma$ orbit of any point is locally finite. Thus the set $\\mathcal{B}$ is finite. Therefore, the geodesic-covering number of $F_T$ in $\\Gamma$ is finite. The same proof also works for the closure of $F_T$. \n\\end{proof}\n\n\\begin{remark}\n\tExplicitly counting the cardinality of a set of the type $\\mathcal{B}$ is the so-called \\textit{hyperbolic circle problem}, see e.g. Lax-Phillips \\cite{Lax-Phillips} and Phillips-Rudnick \\cite{Phillips-Rudnick} etc.\n\\end{remark}\n\n\n\n\\section{Finite index subgroups of the modular group}\nIn this section, we give the proof of Theorem \\ref{thm-finite-index}. \n\nLet $\\Gamma$ be a finite index subgroup of $\\mathrm{PSL}(2, \\mathbb{Z})$ with $[\\mathrm{PSL}(2,\\mathbb{Z}):\\Gamma]=\\mu$. Let $F$ be a fundamental domain of $\\mathrm{PSL}(2, \\mathbb{Z})$, we may choose \n$$F=\\Big\\{z\\in\\mathbb{H}^2: |\\Re(z)|\\leq \\frac{1}{2}, |z|\\geq 1 \\Big\\}.$$ \nIf we have the right coset decomposition $$\\mathrm{PSL}(2,\\mathbb{Z})=\\bigcup_{i=1}^{\\mu} \\Gamma \\alpha_i,$$ then \n\\begin{equation}\\label{finite-index-fundamental-domain}\nF_{\\Gamma}=\\bigcup_{i=1}^{\\mu} \\alpha_i(F)\n\\end{equation}\n is a fundamental domain of $\\Gamma$. One can choose the coset representatives properly to get a simply connected fundamental domain of $\\Gamma$ (see \\cite{Schoen}, Chapter IV, Theorem 3).\nFor example, for the principal congruence subgroup $$\\Gamma(2)=\\left\\{ \\gamma\\in\\mathrm{PSL}(2, \\mathbb{Z}): \\gamma\\equiv \\begin{pmatrix}\n1 & 0\\\\\n0 & 1\n\\end{pmatrix} \\bmod 2 \\right\\},$$\n with index $[\\mathrm{PSL}(2, \\mathbb{Z}):\\Gamma(2)]=6$, see Figure \\ref{figure-Gamma2} (the arrows show the side parings) for a fundamental domain of $\\Gamma(2)$ and Figure \\ref{figure-surface-Gamma2} the shape of the surface $\\Gamma(2)\\backslash\\mathbb{H}^2$. \n\n\n\n\\begin{figure}[ht]\n\t\\centering\n\t\\includegraphics[width=0.5\\textwidth]{fundamental-domain-Gamma2.jpg}\n\\caption{Fundamental domain for $\\Gamma(2) $}\n\t\\label{figure-Gamma2}\n\t\\end{figure}\n\\begin{figure}[ht]\n\t\\centering\n\t\\includegraphics[width=0.7\\textwidth]{Gamma2-surface.png}\n\t\\caption{Shape of surface $\\Gamma(2)\\backslash\\mathbb{H}^2$}\n\t\\label{figure-surface-Gamma2}\n\\end{figure}\n\n\n\n\n\nFor a set $\\mathcal{S}$ of $N$ points on $Y\\cong\\Gamma\\backslash\\mathbb{H}^2$, we consider their representatives in a fundamental domain $F_\\Gamma$ constructed from the right coset decomposition. Since $F_{\\Gamma}$ is a union of $\\mu$ copies of $F$, there exists an $\\alpha_j$ such that $\\alpha_j(F)$ contains $\\geq N\/\\mu$ points from $\\mathcal{S}$. Without loss of generality, we may assume $\\alpha_j$ is identity and still denote this copy as $F$. Otherwise, we just take $\\alpha_j^{-1}(F_{\\Gamma})$ as the fundamental domain of $\\Gamma$ since $\\alpha_j$ is an isometry of $\\mathbb{H}^2$ and this transformation will not change distances and angles among the points we are considering. If we have a lower bound for distinct distances among these $\\geq N\/\\mu$ points, this would also give us a lower bound for distinct distances among all points of $\\mathcal{S}$. \n\nWe divide $F$ into two parts $F=F_{u}\\cup F_o$ (see Figure \\ref{figure-partition-fundamental-domain}) with \n\\begin{equation}\\label{one-fundamental-domain-parition}\nF_u:=\\Big\\{ z=x+iy\\in\\mathbb{H}^2: |x|\\leq \\frac{1}{2}, y\\geq 2 \\Big\\} ~\\text{and}~\nF_o:=F\\setminus F_u.\n\\end{equation}\n\\begin{figure}[ht]\n\t\\centering\n\t\\includegraphics[width=0.6\\textwidth]{fundamental-domain-partition.jpg}\n\t\\caption{Partition of the fundamental domain $F=F_u+F_o$}\n\t\\label{figure-partition-fundamental-domain}\n\\end{figure}\n\n\nWe want to bound the geodesic-covering numbers of $F_u$ and $F_o$ in different base groups. We prove the following lemma. \n\\begin{lemma}\\label{lem-finite-index-geo-cover-number}\nFor any subgroup $\\Gamma$ of $\\mathrm{PSL}(2, \\mathbb{Z})$, the geodesic-covering numbers $K_{\\Gamma}(F_u)$ and $K_{\\Gamma}(F_o)$ are both bounded by some absolute constants. Precisely\t\n\\begin{enumerate}[label=(\\roman*)]\n\t\\item\\label{finite-index-cusp-part} The geodesic-covering number of $F_u$ in $\\Gamma$ is $K_{\\Gamma}(F_u)\\leq 3$.\n\t\\item\\label{finite-index-central-part} The geodesic-covering number of $F_o$ in $\\Gamma$ is $K_{\\Gamma}(F_o)\\leq 252$.\n\\end{enumerate}\n\\end{lemma}\n\\begin{remark}\n\tThe estimate in \\ref{finite-index-central-part} may be improved by more careful calculations. We don't aim to optimize the constant here. The key point is that the geodesic-covering number of $F_o$ in any subgroup $\\Gamma$ is absolutely bounded and thus independent of the index of $\\Gamma$ in $\\mathrm{PSL}(2, \\mathbb{Z})$. One may also use $y\\geq U$ in the definition of $F_u$ for any large enough $U$ to optimize the estimate of $K_{\\Gamma}(F_o)$. \n\\end{remark}\n\n\n\n\nBefore giving the proof of Lemma \\ref{lem-finite-index-geo-cover-number}, we use it to prove Theorem \\ref{thm-finite-index} first. \n\\begin{proof}[Proof of Theorem \\ref{thm-finite-index}]\nSuppose $\\Gamma$ is a subgroup of $\\mathrm{PSL}(2, \\mathbb{Z})$ of finite index $[\\mathrm{PSL}(2,\\mathbb{Z}):\\Gamma]=\\mu$. \t\nLet $\\mathcal{S}$ be a set of $N$ points on the hyperbolic surface $Y\\cong\\Gamma\\backslash\\mathbb{H}^2$, and define the distance set\n$$d_Y(\\mathcal{S}):=\\{ d_Y(p, q): p, q\\in \\mathcal{S} \\}.$$\nIf $F$ is a fundamental domain of $\\mathrm{PSL}(2, \\mathbb{Z})$, by the fundamental domain of $\\Gamma$ in the form \\eqref{finite-index-fundamental-domain}, there exists some $j$ such that $\\alpha_j(F)$ contains more than $N\/\\mu$ points. Since $\\alpha_j$ is an isometry of $\\mathbb{H}^2$, without loss of generality, we assume $\\alpha_j(F)=F$ and let $\\mathcal{S}_F$ be these $\\geq N\/\\mu$ points on it. We observe that \n\\begin{equation}\\label{eq-distance-surface-to-one-F}\n|d_Y(\\mathcal{S})|\\geq |d_Y( \\mathcal{S}_F )|. \n\\end{equation}\nWe use Lemma \\ref{lem-finite-index-geo-cover-number} to establish a lower bound for $\t|d_Y(\\mathcal{S}_F )|$ and hence derive a lower bound for $|d_Y(\\mathcal{S})|$. \n\nWe partition the region $F=F_u\\cup F_o$ as in \\eqref{one-fundamental-domain-parition}. Either $F_u$ or $F_o$ contains more than $\\frac{1}{2}|\\mathcal{S}_F|\\geq N\/2\\mu$ points. \n\n\\textbf{Case 1).} The region $F_u$ contains more than $\\frac{1}{2}|\\mathcal{S}_F|$ points. Let $\\mathcal{S}_u:=\\mathcal{S}_F\\cap F_u$ be the points on $F_u$, and $\\Gamma_u$ be a geodesic-cover of $F_u$ in $\\Gamma$ with cardinality $K_{\\Gamma}(F_u)$. Then we have\n\\begin{align}\nQ_Y(\\mathcal{S}_u )&:=\\big\\{(p_1, p_2; p_3, p_4)\\in \\mathcal{S}_u^4: d_Y(p_1, p_2)=d_Y(p_3, p_4)\\neq 0 ) \\big\\}\\nonumber\\\\\n&~\\subset Q_{\\mathbb{H}^2}\\big(\\cup_{\\gamma\\in\\Gamma_u} \\gamma(\\mathcal{S}_u) \\big),\n\\end{align}\nwhere $Q_{\\mathbb{H}^2}(\\mathcal{P})$ is defined in \\eqref{eq-defn-QH}. By Lemma \\ref{lem-finite-index-geo-cover-number} \\ref{finite-index-cusp-part} and \\eqref{eq-Tao-quaruple}, we derive \n\\begin{equation}\n|Q_Y(S_u)|\\ll K_{\\Gamma}^3(F_u) |\\mathcal{S}_u|^3 \\log(K_{\\Gamma}(F_u)|\\mathcal{S}_u|)\\leq 27 |\\mathcal{S}_u|^3 \\log(3|\\mathcal{S}_u|).\n\\end{equation}\nConsequently by \\eqref{Quadruple-to-distance} and \\eqref{eq-distance-surface-to-one-F}, we get the lower bound\n\\begin{equation}\n|d_Y(\\mathcal{S})|\\geq |d_Y(\\mathcal{S}_u)|\\gg \\frac{|\\mathcal{S}_u|}{ \\log|\\mathcal{S}_u| },\n\\end{equation}\nwhere the implied constant is absolute. Therefore, by the assumption $\\frac{N}{2\\mu}\\leq |\\mathcal{S}_u|\\leq N$, we conclude that\n\\begin{equation}\\label{eq-distance-finite-index-cuspidal}\n|d_Y(\\mathcal{S})|\\geq C_1\\frac{N}{\\mu\\log N}\n\\end{equation}\nfor some absolute constant \t$C_1>0$. \n\t\n\t\n\\textbf{Case 2).}\tThe region $F_o$ contains more than $\\frac{1}{2}|\\mathcal{S}_F|$ points. Let $\\mathcal{S}_o=\\mathcal{S}_F\\cap F_o$ and $\\Gamma_o$ be a geodesic cover of $F_o$ in $\\Gamma$ with cardinality $K_{\\Gamma}(F_o)$. By Lemma \\ref{lem-finite-index-geo-cover-number} \\ref{finite-index-central-part} and a similar argument as in \\textbf{Case 1)}, we derive that\n\\begin{equation}\nQ_Y(\\mathcal{S}_o)\\subset Q_{\\mathbb{H}^2}\\big( \\cup_{\\gamma\\in\\Gamma_o}\\gamma(F_o) \\big)\n\\end{equation}\nand thus\n\\begin{equation}\n|Q_Y(\\mathcal{S}_o)|\\ll K_{\\Gamma}^3(F_o) |\\mathcal{S}_o|^3 \\log(K_{\\Gamma}(F_o)|\\mathcal{S}_o|)\\leq 252^3 |\\mathcal{S}_o|^3 \\log( 252 |\\mathcal{S}_o|). \n\\end{equation}\t\nAgain by \\eqref{Quadruple-to-distance} and the assumption $\\frac{N}{2\\mu}\\leq |\\mathcal{S}_o|\\leq N$, we conclude that\n\\begin{equation}\\label{eq-distance-finite-index-central}\n|d_Y(\\mathcal{S})|\\geq |d_Y(\\mathcal{S}_o)|\\geq C_2\\frac{N}{\\mu\\log N}\n\\end{equation}\t\nfor some absolute constant $C_2>0$. \n\nFinally, combining \\eqref{eq-distance-finite-index-cuspidal} and \\eqref{eq-distance-finite-index-central} and taking $C=\\min\\{C_1, C_2\\}$, we get the desired lower bound for distinct distances in hyperbolic surfaces associated with any finite index subgroup of $\\mathrm{PSL}(2, \\mathbb{Z})$, \n\\begin{equation}\n|d_Y(\\mathcal{S})|\\geq C\\frac{N}{\\mu\\log N}\n\\end{equation}\nfor some absolute constant $C>0$. \n\\end{proof}\n\nIn the following we prove Lemma \\ref{lem-finite-index-geo-cover-number}.\n\n\\begin{proof}[Proof of \\ref{finite-index-cusp-part} in Lemma \\ref{lem-finite-index-geo-cover-number}]\nRecall that $F_u$ is the region \n$$\\Big\\{ z=x+iy\\in\\mathbb{H}^2: |x|\\leq \\frac{1}{2}, y\\geq 2 \\Big\\}.$$\nWe consider the set\n\\begin{equation}\\label{eq-set-smaller-distance}\n\\mathcal{A}:=\\{\\gamma\\in \\mathrm{PSL}(2, \\mathbb{Z}): d_{\\mathbb{H}^2}(z_1, \\gamma z_2)\\leq d_{\\mathbb{H}^2}(z_1, z_2), z_1, z_2\\in F_u, {\\rm Im}(z_1)\\geq {\\rm Im}(z_2) \\},\n\\end{equation} \nwhich is a geodesic cover of $F_u$ in $\\mathrm{PSL}(2, \\mathbb{Z})$ by Definition \\ref{defn-geodesic-covering-number}. For any subgroup $\\Gamma$ of $\\mathrm{PSL}(2, \\mathbb{Z})$, the set $\\mathcal{A}\\cap\\Gamma$ is a geodesic cover of $F_u$ in $\\Gamma$. \n\nFor any two points $z_1=x_1+iy_1$ and $z_2=x_2+iy_2$ in $F_u$ with $y_1\\geq y_2\\geq 2$, and\n $$\\gamma=\\begin{pmatrix}\n a&b\\\\\n c&d\n \\end{pmatrix}\\in \\mathrm{PSL}(2,\\mathbb{Z}),$$\nthe imaginary part of $\\gamma(z_2)$ can be written as\n$$ \\frac{y_2}{|cz_2+d|^2}=\\frac{y_2}{(cx_2+d)^2+c^2 y_2^2 }.$$\n\nIf $c=0$, then $a=d=1$, the isometry $\\gamma$ is actually a translation of the form\n$$\\gamma=\\begin{pmatrix}\n1&b\\\\\n0&1\n\\end{pmatrix}\\in \\mathrm{PSL}(2, \\mathbb{Z}),$$\nfor some $b\\in\\mathbb{Z}$. The only possible choices of $\\gamma$ for which $d_{\\mathbb{H}^2}(z_1, \\gamma z_2)\\leq d_{\\mathbb{H}^2}(z_1, z_2)$ are from the set\n\\begin{equation}\\label{eq-translation-cover}\n\\mathcal{T}=\\left\\{ \\begin{pmatrix}\n1 & -1\\\\\n0 & 1\n\\end{pmatrix}, \n\\begin{pmatrix}\n1 &0\\\\\n0 & 1\n\\end{pmatrix},\n\\begin{pmatrix}\n1 & 1\\\\\n0 & 1\n\\end{pmatrix} \\right\\}.\n\\end{equation} \n\nIf $c\\neq 0$, then $|c|\\geq 1$ and thus\n$${\\rm Im}(\\gamma z_2)\\leq \\frac{1}{y_2}\\leq \\frac{1}{2}.$$\nDenote $\\gamma(z_2)=x_0+iy_0$, then $y_0\\leq \\frac{1}{2}$. By the hyperbolic distance formula \\eqref{eq-hyperbolic-cosh-distance}\nwith the fact $y_1\\geq y_2\\geq 2$ and $|x_1-x_2|\\leq 1$, we get\n\\begin{align}\n&2\\cosh(d_{\\mathbb{H}^2}(z_1, \\gamma z_2))\n-2\\cosh(d_{\\mathbb{H}^2}(z_1, z_2)) \\nonumber\\\\\n&=\\frac{(x_1-x_0)^2+y_1^2+y_0^2}{y_1 y_0 }- \\frac{(x_1-x_2)^2+y_1^2+y_2^2}{ y_1 y_2} \\nonumber\\\\\n&\\geq \\frac{y_1}{y_0}-\\frac{1}{y_1 y_2}-\\frac{y_1}{y_2}-\\frac{y_2}{y_1} \\nonumber\\\\\n&\\geq 2y_1-\\frac{1}{4}-\\frac{y_1}{2}-1\\geq \\frac{7}{4}>0.\n\\end{align}\nThus for any $\\gamma\\in \\mathrm{PSL}(2, \\mathbb{Z})$ with $c\\neq 0$, we always have $d_{\\mathbb{H}^2}(z_1, \\gamma z_2)>d_{\\mathbb{H}^2}(z_1, z_2)$. Hence $\\mathcal{A}=\\mathcal{T}$. \n\nFor $\\Gamma$ being any subgroup of $\\mathrm{PSL}(2, \\mathbb{Z})$, the elements of $\\gamma'\\in\\Gamma$ such that $$d_{\\mathbb{H}^2}(z_1, \\gamma' z_2)\\leq d_{\\mathbb{H}^2}(z_1, z_2) ~\\text{ with} z_1, z_2\\in F_u, {\\rm Im}(z_1)\\geq {\\rm Im}(z_2)$$ \nare also from the set $\\mathcal{A}=\\mathcal{T}$ in \\eqref{eq-set-smaller-distance} and \\eqref{eq-translation-cover}. Therefore, by Definition \\ref{defn-geodesic-covering-number}, the set $\\mathcal{T}\\cap \\Gamma$ is a geodesic cover of $F_u$ in $\\Gamma$. (Note that $\\mathcal{T}\\cap\\Gamma$ always contains the identity.) We conclude that the geodesic-covering number of $F_u$ in any subgroup $\\Gamma$ of $\\mathrm{PSL}(2, \\mathbb{Z})$ is $K_{\\Gamma}(F_u)\\leq 3$. \n\\end{proof}\n\n\n\\begin{proof}[Proof of \\ref{finite-index-central-part} in Lemma \\ref{lem-finite-index-geo-cover-number}]\nNow we deal with the bounded part $$F_o=\\{z=x+iy\\in\\mathbb{H}^2: |z|\\geq 1, 02$ iterations and thus the numerical search for the optimal gate, i.e, \nthe search for $\\boldsymbol{\\alpha}^* \\in \\mathbb{R}^{15}$, becomes inefficient. \n\n\\begin{figure*}[t!]\n \\begin{center}\n \\includegraphics[width=.39\\textwidth]{fig3a.pdf}\n \\includegraphics[width=.39\\textwidth]{fig3b.pdf}\n\\caption{Optimized purification protocol for Werner states, see Eq.~\\eqref{eq:wernerstate}.\nLeft panel: average cost function $\\bar{f}$ with a uniform PDF as a function of $N$, the number of iterations. Right panel: The overall success \nprobability after $N=1$ (dotted line), $N=2$ (dash-dotted line), and $N=3$ (full line) iterations as a function of $x$.}\n \\label{fig:Werner}\n \\end{center}\n\\end{figure*}\n\nInspired by {\\it Example 2.}, we consider the state in Eq.~\\eqref{eq:example2} and also\n\\begin{equation}\nx\\ketbra{\\Psi^-}{\\Psi^-}+(1-x)\\ketbra{\\Upsilon}{\\Upsilon} \\label{eq:MaZs}\n\\end{equation}\nwith $x \\in [0,1]$ and \n\\begin{equation}\n \\ket{\\Upsilon}=\\frac{1}{\\sqrt2} \\left(\\ket{\\Psi^+}+i\\ket{\\Phi^-}\\right). \\nonumber\n\\end{equation}\nBy using the unitary transformation in Eq.~\\eqref{eq:localtr} on Eq.~\\eqref{eq:MaZs} one obtains Eq.~\\eqref{eq:example2} and therefore it is expected that the only optimal two-qubit gates are the ones that purify these states\nin one iteration. Both states have an input concurrence $\\mathcal{C}(x)=x$. However, it is important to note that the CNOT gate is not optimal for the state in Eq.~\\eqref{eq:MaZs}, because after one iteration round\n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x)=\\begin{cases} 0& x \\in [0,0.5], \\\\ 2x \\frac{2x-1}{1+x^2} & x \\in (0.5,1], \\end{cases} \\nonumber\n\\end{equation} \nand $\\mathcal{C}'_{\\text{CNOT}}(x) \\leqslant x$. The success probability is\n\\begin{equation}\n P_{\\text{CNOT}}=\\frac{1+x^2}{2}. \\nonumber\n\\end{equation}\nThus, the CNOT gate based purification protocol impairs the concurrence. In contrast to this analytical observation, numerical analysis with uniform PDF yields already in the first iteration for both states an average cost function $\\bar{f}\\approx 0.0002$. To demonstrate the robustness of the numerical approach we provide examples of three non-uniform PDFs for $N=1$ \niteration. First, we consider\n\\begin{equation}\np(x)=2x,\\quad \\text{with} \\quad x \\in [0,1], \\nonumber\n\\end{equation}\nwhich describes a situation, where states with higher concurrences are more likely to be subject to the purification. The resulting average cost function is $\\bar{f}\\approx 0.000004$. Secondly, we take \n\\begin{equation}\np(x)=2(1-x),\\quad \\text{with} \\quad x \\in [0,1], \\nonumber\n\\end{equation}\nwhich puts more weight on states with low concurrences and find $\\bar{f}\\approx 0.0005$. Finally, we investigate a PDF\n\\begin{equation}\np(x)=6x(1-x),\\quad \\text{with} \\quad x \\in [0,1], \\nonumber\n\\end{equation}\ni.e., the states around the concurrence $\\mathcal{C}(x)=0.5$ are more likely to participate in the purification, and obtain $\\bar{f}\\approx 0.00005$. These results demonstrate the effectiveness of our approach and up to a numerical precision these one-parameter families of states can be purified in one iteration.\n\n\\begin{figure*}[t!]\n \\begin{center}\n \\includegraphics[width=.39\\textwidth]{fig4a.pdf}\n \\includegraphics[width=.39\\textwidth]{fig4b.pdf}\n\\caption{Optimized purification protocol for the state in Eq.~ \\eqref{eq:qrstate}. Left panel: Average cost function $\\bar{f}$ with a uniform PDF as a function of $N$, the number of iterations. \nCrosses are the results of the CNOT gate, whereas dots display the numerical optimization. They have been connected by lines to guide the eye. Right panel: The concurrence $\\mathcal{C}'$\nas a function of $x$ and $y$ after one purification round. The cone-type surface is obtained with our approach. A less optimal surface with minima at $x=0$ and along the $y$ axis is the result of the purification protocol \nwith the CNOT gate.}\n \\label{fig:XY}\n \\end{center}\n\\end{figure*}\n\nNext, we consider the following two-parameter family of states \\cite{Bernad2}\n\\begin{equation}\n\\hat{\\rho}(x,y) = \\frac{1}{4}\n\\left(\n\\begin{array}{cccc}\n 1-x & i y & -i y & x-1 \\\\\n -i y & x+1 & -x-1 & i y \\\\\n i y & -x-1 & x+1 & -i y \\\\\n x-1 & -i y & i y & 1-x \\\\\n\\end{array}\n\\right) \\label{eq:qrstate}\n\\end{equation}\nwith $x,y \\in \\mathbb{R}$ and $x^2+y^2\\leqslant1$. The concurrence of this state is $\\sqrt{x^2+y^2}$. In order to relate the performance of our approach, we apply the CNOT gate based purification \nprotocol to this family of states. One obtains\n\\begin{eqnarray}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)&=&\\frac{2|x|}{1+x^2}, \\label{eq:xyCNOTc} \\\\\n P_{\\text{CNOT}}&=&\\frac{1+x^2}{2}, \\label{eq:xyCNOTp}\n\\end{eqnarray}\nafter $N=1$ and\n\\begin{eqnarray}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)&=&\\frac{4 |x| (1+x^2)}{1+6x^2+x^4}, \\label{eq:xyCNOTc2} \\\\\n P_{\\text{CNOT}}&=&\\frac{1+6x^2+x^4}{2 (1+x^2)^2}, \\label{eq:xyCNOTp2}\n\\end{eqnarray}\nafter $N=2$ purification rounds, respectively. If one applies the unitary transformation in Eq.~\\eqref{eq:localtr} on the state before the bilateral CNOT gates are performed, then the above results remain \nunchanged. These results demonstrate that the CNOT gate and the additional tricks, which yielded optimal purifications for the one-parameter family of states, are not optimal in this scenario, since they are outperformed on average by our optimized protocol. \nWe assume that CNOT-based purification protocols cannot exploit all the useful entanglement, because the concurrence \nand the success probability become independent of the $y$ variable. Using these analytical results and the uniform PDF\n\\begin{equation}\n p(x,y)=\\frac{1}{\\pi},\\quad \\text{with} \\quad x^2+y^2 \\leqslant 1, \\nonumber\n\\end{equation}\nwe compare our approach with the above-presented analytical results, see Eq.~\\eqref{eq:xyCNOTc} and Eq.~\\eqref{eq:xyCNOTc2}. In Fig.~\\ref{fig:XY}, we show that our optimized protocol provides after one iteration a $x$ and $y$-dependent concurrence. Although the \nnumerically-obtained concurrence is larger at $x=0$ and $y \\in [-1,1]$ than the surface given by Eq.~\\eqref{eq:xyCNOTc}, one can also observe the opposite at $y=0$ and $x \\in [-1,1]$. However, the concurrence presents a better improvement with\nour algorithm and the average cost function with the uniform PDF stays for more iterations lower than the CNOT-based protocol, see the left figure in panel \\ref{fig:XY}. \nIn regard to the overall success probability, we let our algorithm work until the third iteration and in Fig.~\\ref{fig:XYos} the results are compared with the CNOT-based purification protocol. The obtained \nsurfaces differ only by a $\\pi\/2$ rotation around the $z$-axis. Therefore, there is not much difference in the overall success probability and from this aspect the CNOT gate can also be considered optimal, though, \nit still can not improve the concurrence as effectively as the gate obtained with our approach. However, one might think with a proper choice of local unitary transformations, like the transformation \nin Eq.~\\eqref{eq:localtr} used for Werner and one-step purifiable states, the application of the CNOT gate may result in an optimal purification protocol. Let us consider then two general local unitary \ntransformations at locations $A$ and $B$ for the preparation of two-qubit states. It is enough to consider unitary transformations in $SU(2)$, for which the Euler angle parameterization reads \\cite{Tilma2}\n\\begin{eqnarray}\n U_A&=& e^{i \\sigma_z \\alpha_1} e^{i \\sigma_y \\alpha_2} e^{i \\sigma_z \\alpha_3}, \\\\\n U_B&=& e^{i \\sigma_z \\beta_1} e^{i \\sigma_y \\beta_2} e^{i \\sigma_z \\beta_3},\n\\end{eqnarray}\nwhere $\\sigma_y$ and $\\sigma_z$ are the Pauli matrices. Here, $\\alpha_1, \\beta_1 \\in [0, \\pi]$, $\\alpha_2,\\beta_2 \\in [0, \\pi\/2]$, and $\\alpha_3, \\beta_3 \\in [0, 2\\pi]$ are the Euler angles for $SU(2)$. For the first\niteration, we analyze the output\nunitary of the optimized protocol with \\textit{Mathematica} \\cite{CAS} and observe that the angles $\\alpha_1$ and $\\beta_1$ do not affect the efficiency of the CNOT-based protocol. Optimal \nconcurrences for the remaining four angles yield either the result in Eq.~\\eqref{eq:xyCNOTc} or\n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)=\\frac{2|y|}{1+y^2}, \n\\end{equation}\nwith success probability\n\\begin{equation} \n P_{\\text{CNOT}}=\\frac{1+y^2}{2}, \n\\end{equation}\nfor $\\alpha_2=\\frac{\\pi}{8}$, $\\beta_2=\\frac{3 \\pi}{8}$, and $\\alpha_3=\\beta_3=\\frac{3\\pi}{4}$. For these results the average cost function with uniform PDF yields $\\bar{f}\\approx 0.372$, which is still lower than\nthe average concurrence found by our approach. These results demonstrate that even with local unitary transformations the CNOT gate is globally not optimal for all values of $x$ and $y$. However, when $x$ and $y$\nare known beforehand, then one can use the following local unitary transformation\n\\begin{eqnarray}\n U^\\dagger_A \\otimes U_B, \\quad \\text{with} \\quad U=\\cos(\\theta)\\mathbb{I}_2 + \\sin(\\theta) \\sigma_x, \\label{eq:statedeptr}\n\\end{eqnarray}\nwhere the angle $\\theta$ is a function of $x$ and $y$. For example, for $x,y \\geqslant 0$, we have\n\\begin{equation}\n \\cos(\\theta)=\\sqrt{\\frac{1}{2}+\\sqrt{\\frac{x+\\sqrt{x^2+y^2}}{8 \\sqrt{x^2+y^2}}}}. \\nonumber\n\\end{equation}\nUsing this transformation before the purification protocol, one obtains the state\n\\begin{equation}\n\\frac{1+ \\sqrt{x^2+y^2}}{2} \\ketbra{\\Psi^-}{\\Psi^-}+\\frac{1- \\sqrt{x^2+y^2}}{2} \\ketbra{\\Phi^-}{\\Phi^-}. \\label{eq:Spec}\n\\end{equation}\nNow, the CNOT-based purification yields\n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)=\\frac{2\\sqrt{x^2+y^2}}{1+x^2+y^2}, \n\\end{equation}\nwith success probability\n\\begin{equation}\n P_{\\text{CNOT}}=\\frac{1+x^2+y^2}{2}. \n\\end{equation}\nIt is immediate that $\\mathcal{C}'_{\\text{CNOT}}(x,y) \\geqslant \\sqrt{x^2+y^2}$, i.e., we are improving the concurrence. In this particular case, one does not need the average cost function in the numerical search, \nbecause the state is fixed. Our algorithm running with a single state with parameters $(x,y)$ as defined in Eq.~\\eqref{eq:qrstate} instead of an ensemble of states can improve the concurrence, but is unable to find the optimal solution presented above, because the transformation\nEq.~\\eqref{eq:statedeptr} together with CNOT gates is a non-symmetrical transformation at nodes $A$ and $B$.\n\nFinally, let us point out that our approach does not take into account operational or memory errors. These always depend on the implementation and if an experiment can provide us models for the errors then \nour approach can be easily extended. Another important experimental input is the PDF $p(\\boldsymbol{x})$, which is usually subject to the method of generating entangled states between locations A and B. Furthermore, \nthis PDF is assigned to the process of choosing a value $\\boldsymbol{x}$ as a random event, i.e, we have the same two two-qubit pairs parameterized by $\\boldsymbol{x}$ before the purification protocol. In effect, \nwe integrate the parameter dependence of the states away and thus our approach yields an optimal two-qubit gate on average. If the value of $\\boldsymbol{x}$ is fixed in an experimental design, then our method \nprovides an optimal two-qubit gate designed for this particular state. The optimal two-qubit gates can be can be realized by three\nCNOT gates and additional single-qubit rotations \\cite{Vidal,Vatan}, and therefore the experimental generation of this gate is possible with high fidelity \\cite{Debnath}. Also in the context of trapped ions or superconducting quantum circuits, the generation of two-qubit entangling gates can be achieved with high precision using the M\\o{}lmer-S\\o{}rensen gate \\cite{MS} or the $\\sqrt{i \\text{SWAP}}$ \\cite{Fan} gate. Since two-qubit unitaries have a straightforward implementation in many physical settings, thanks to quantum compilation \\cite{Martinez}, we argue that quantum computing platforms implementing quantum communication could actually benefit from globally optimal gates for entanglement purification.\n\n\\begin{figure*}[t!]\n \\begin{center}\n \\includegraphics[width=.39\\textwidth]{fig5.pdf}\n\\caption{The overall success probability of the state in Eq.~ \\eqref{eq:qrstate} after and $N=3$ iterations as a function of $x$ and $y$. Both surfaces are similar in appearance. The one having maxima at $x=1$ \nbelongs to the CNOT-based protocol and the other one having maxima at $y=1$ is obtained via the optimized numerical protocol.}\n \\label{fig:XYos}\n \\end{center}\n\\end{figure*}\n\n\\section{Conclusions}\n\\label{Conclusions}\n\nIn the context of entanglement purification and recurrence protocols, we have presented a method to obtain optimal protocols. This method searches for the optimal two-qubit gate, which is applied bilaterally\nat the nodes $A$ and $B$ in order to distill from an ensemble of noisy pairs a higher fidelity state with respect to a maximally entangled state. Our strategy considers situations, where the \nexperimental apparatus does not have enough control over the entanglement generation phase and different experimental runs result in different states. Here, we assume that the same copies of the noisy state can be \ngenerated before the purification protocol takes place, but a different experimental run could result in different states. Errors originating from local operations, memory requirements, or even \nclassical communication are neglected for now.\n\nWe have numerically demonstrated the optimality of our proposal for several states. In the case of the Werner state, we have found several optimal two-qubit gates and their performances are the same as the CNOT-based Bennett protocol \\cite{Bennett}. Thus, for Werner states the optimality cannot be improved beyond the already known performance. Next, we have investigated a family of states that one can purify \nin one step, i.e., two noisy copies of states are enough to obtain a maximally entangled state, where we immediately obtain minimal average cost function $\\bar{f} \\approx 0$, as expected. Finally, we consider a two-parameter family of states which is considered in theoretical models of a quantum repeater \\cite{Bernad2}. Our numerical investigations demonstrate that a\nsingle bilaterally applied CNOT gate cannot be globally optimal for these states. On the other hand, when the state is known, then parameter-dependent, non-symmetric local transformations allow again the CNOT to be also \none of the optimal two-qubit gates. This case exemplifies the difference between protocols that are optimal for a full class of parameterized states and those that are only optimal for a single state. \n\nIn conclusion, we have proposed a general method to optimize entanglement purification for an arbitrary family of states. Our algorithm \\cite{Francesco} can find the most optimal two-qubit gate that on average induces the highest increase of entanglement. We remark that the method is general for the set of parameters defining the states. Optimizing for a single state is also possible, as shown in one of the \npresented examples. Furthermore, we have focused on the degree of entanglement measured by the concurrence and not on the fidelity with respect to a particular Bell state. \nThis is motivated by the fact that there exist an infinite number of maximally entangled states, and therefore entangled mixed states can present a larger fidelity with respect to maximally entangled states that might not correspond to Bell states defined in the computational basis. \nA possible drawback is that one cannot know with certainty the final maximally entangled state produced by the protocol. This and other issues such as different input states, asymmetric two-qubit gates, \nand non-ideal local operations remain open questions. However, this work aims to introduce the concept of globally optimized recurrence protocols that is flexible enough to incorporate the \naforementioned points in further investigations, which we plan to carry out in the future. \n\n\\begin{acknowledgments}\nThis work is supported by the DFG under Germany's Excellence Strategy-Cluster of Excellence Matter and Light for Quantum\nComputing (ML4Q) EXC 2004\/1-390534769. The results were obtained using the libraries JAX \\cite{JAX}, Scipy \\cite{Scipy} and QuTiP \\cite{Qutip}.\n\n\\end{acknowledgments}\n\n\n\\section{Introduction}\n\nEntanglement is a key resource for quantum information tasks, like quantum simulations \\cite{Monroe}, computation \\cite{Albash}, and communication \\cite{Pirandola}. These tasks are based on the idea of creating networks of quantum systems, where the generation of well-controlled entanglement between spatially separated nodes is essential. These networks, which may consist of distant or nearby nodes, have been thoroughly investigated for efficient process and transfer of quantum information \\cite{Awschalom}. However, due to interactions with an uncontrollable environment noisy or non-maximally entangled states are produced. To protect quantum information and to guarantee a high performance of its processing, one can use quantum error correction \\cite{Peres, Devitt} or quantum teleportation in combination with entanglement purification \\cite{Bennett1, Deutsch, Bennett2}. Quantum error correction is particularly suitable for quantum computers, however it is inferior to entanglement purification with two-way classical communication \\cite{Bennett2}, which is the subject of our investigations. In particular, we consider a so-called recurrence protocol \\cite{Dur}, an iterative approach, which operates in each purification step only on two identical copies of states.\n\nIn this paper, we discuss entanglement purification from the point of view of optimality. Recently, optimized entanglement purification has been investigated with the help of genetic algorithms \\cite{Krastanov}, where \nthe analytical and numerical studies are based on Werner states \\cite{Werner}. In fact, also the first ever proposed protocols are based on either Werner \\cite{Bennett1} or Bell diagonal states \\cite{Deutsch}. It has been \nshown that $4$ or $12$ local random $SU(2)$ transformations can bring any states into a Bell diagonal or Werner state, respectively \\cite{Bennett2}. We have already argued that these random local unitary transformations \nand states obtained in consequence are only useful\nfor grasping and consequently understanding the complex task of purification because they may waste useful entanglement \\cite{Torres}. Therefore, here, we develop a method for general states and demonstrate it for \nseveral simple examples including also the Werner state. Nonetheless, our motivation also lies in the fact that an experimental implementation may not have enough control over the generated noisy states and thus more general, adaptive, and optimal purification strategies have to be made available. In this general context, we assume that an experiment can still guarantee identical copies of states before the purification protocol takes place. \n\nOur method is based on quasi-Monte Carlo numerical sampling of the states, which undergo the purification protocol, and then the concurrence \\cite{Wooters} of the output states is integrated over an {\\it a priori} probability distribution \nfunction. This results in an average two-qubit gate-dependent cost function for the whole sample of states and then we employ a quasi-Newton algorithm \\cite{Kelley} for solving this non-linear optimization problem \non the whole $SU(4)$ group. The obtained two-qubit gate is optimal in the sense that it achieves, on average, a higher increase of entanglement of an input family of states. We discuss the performance of our method for several examples and compare with protocols based on one bilateral application of controlled-NOT gates, the paradigmatic two-qubit operation used in the seminal purification protocols \\cite{Bennett1,Deutsch}.\n\nThe paper is organized as follows. In Sec.~\\ref{Method} we introduce our\nmethod and give some elementary examples to allow further acquaintance with the concept of the introduced cost function. \nIn Sec.~\\ref{Results} we demonstrate our approach for four one-parameter and one two-parameter family of states. \nNumerical and analytical results are presented for concurrences and success probabilities. In Sec.~\\ref{Conclusions}\nwe summarize and draw our conclusions. Some details supporting the main text are collected in Appendix~\\ref{AppendixA}.\n\n\\section{Method}\n\\label{Method}\nIn this section, we describe how the recurrence protocol is optimized with respect to an input familiy of quantum states.\n\n\\subsection{Entanglement purification protocol}\\label{EntPurif}\nLet us consider the product state of two-qubit pairs \n\\begin{equation}\n \\boldsymbol{\\rho}=\\rho^{A_1,B_1} \\otimes\\rho^{A_2,B_2},\n\\label{eq:rhoAB4}\n\\end{equation}\nwhere qubit components of each pair are assumed to be spatially separated at distant locations. These locations are labeled by $A$ and $B$. In an \nentanglement purification protocol, one performs local quantum operations, which may not just involve two-qubit gates \\cite{Bernad1}. \nThis is followed by measurements on one of the pairs\nat both locations. A classical communication between $A$ and $B$ results in a qubit pair with a higher degree of entanglement. Both pairs are assumed to start in the same state $\\rho$ \nand the degree of the entanglement is usually measured by the \nfidelity with respect to one of the Bell states\n\\begin{align}\n \\ket{\\Psi^\\pm}=\\tfrac{1}{\\sqrt2}\\left(\\ket{01}\\pm\\ket{10}\\right), \\,\n \\ket{\\Phi^\\pm}=\\tfrac{1}{\\sqrt2}\\left(\\ket{00}\\pm \\ket{11}\\right).\n \\label{eq:Bellstates}\n\\end{align}\nHowever, these states can be subject to local unitary transformations, which can cause some technical difficulties, when one uses fidelity, i.e.~using fidelity as a cost function would force us to search also for additional local unitary operations in order to align the output state of the protocol with the Bell basis. As we intend to \nanalyze the entanglement purification in a very general setup, one requires an entanglement measure that is invariant under local unitary transformations. Therefore, we turn to the concurrence as a measure \nof the the attainability of a maximally entangled state \\cite{Wooters}:\n\\begin{equation}\\label{eq:concurrence}\n \\mathcal{C}(\\rho)=\\max\\{0,\\lambda_1-\\lambda_2-\\lambda_3-\\lambda_4\\}.\n\\end{equation}\nHere $\\lambda_1, \\lambda_2, \\lambda_3, \\lambda_4$ are the square roots of the non-negative eigenvalues of the non-Hermitian matrix\n\\begin{equation}\n \\tilde \\rho = \\rho (\\sigma_{y}\\otimes\\sigma_{y})\\rho^{*}(\\sigma_{y}\\otimes\\sigma_{y}), \\nonumber\n\\end{equation}\nwith $*$ being the complex conjugation in the standard basis and $\\sigma_{y}$ is the Pauli matrix.\n\n\\begin{figure}[t!]\n \\includegraphics[width=.4\\textwidth]{fig1.pdf}\n \\caption{Schematic representation of bipartite entanglement purification, where the entanglement purification protocol trades two \n entangled qubit pairs for a qubit pair with a higher degree of entanglement, which is quantified by the concurrence $\\mathcal{C}$.}\n \\label{fig:purifscheme}\n\\end{figure}\n\nIn this manuscript, we examine and optimize a purification protocol having the following steps:\\\\\n\n(I) Two unitary transformations $\\rho \\rightarrow U(\\boldsymbol{\\alpha}) \\rho U(\\boldsymbol{\\alpha})^\\dagger$ are applied locally at $A$ and $B$ (see Fig.~\\ref{fig:purifscheme}), where $U$ is a general two-qubit unitary described by a parameter vector $\\boldsymbol{\\alpha}$.\nAfter the application of the quantum operation the four-qubit system \nattains the state \n\\begin{equation}\n\\boldsymbol{\\rho}'=\n U_{A1,A2}(\\boldsymbol{\\alpha}) U_{B1,B2}(\\boldsymbol{\\alpha}) \\boldsymbol{\\rho} U^\\dagger_{B1,B2} (\\boldsymbol{\\alpha}) U^\\dagger_{A1,A2} (\\boldsymbol{\\alpha}). \n\\label{eq: maprho}\n\\end{equation}\n\n(II) One of the pairs $(A_2,B_2)$ is then locally measured in the standard basis. There are four possible states, i.e., 4-dimensional vectors, in which one can find the measured pair: \n\\begin{eqnarray}\n \\ket{1}&=&\\ket{00}_{A_2,B_2}, \\quad \\ket{2}=\\ket{01}_{A_2,B_2}, \\nonumber \\\\\n \\ket{3}&=&\\ket{10}_{A_2,B_2}, \\quad \\ket{4}=\\ket{11}_{A_2,B_2}. \\nonumber\n\\end{eqnarray}\nA successful measurement of one of the states $\\ket{i}$ with $i\\in\\{1,2,3,4\\}$ results in\na state for the other qubit\n\\begin{equation}\n \\tilde\\rho^{A_1,B_1}_i= \\frac{\\bra{i} \\boldsymbol{\\rho}' \\ket{i} }{\\mathrm{Tr}\\{\\ketbra{i}{i} \\boldsymbol{\\rho}'\\}}\n\\label{}\n\\end{equation}\nwith probability\n\\begin{equation}\n p_i=\\mathrm{Tr}\\{\\ketbra{i}{i} \\boldsymbol{\\rho}'\\}. \\nonumber\n\\end{equation}\n\n(III) Depending on value of the measurement results, which are communicated between the two parties, the state with largest concurrence and the related success probability are kept, whereas the others are discarded. The output of the protocol is the pair \n$(\\mathcal{C}',P)$, concurrence value $\\mathcal{C}'$ of the state obtained with probability $P$. If there are multiple maxima, e.g., $\\mathcal{C}(\\tilde\\rho^{A_1,B_1}_1)=\\mathcal{C}(\\tilde\\rho^{A_1,B_1}_2)$, then \n$\\mathcal{C}'=\\mathcal{C}(\\tilde\\rho^{A_1,B_1}_1)$ and $P=p_1+p_2$.\\\\\n\nGiven two copies of a state $\\rho$ with concurrence $\\mathcal{C}$, it is straightforward to see that pair $(\\mathcal{C}',P)$ depends on the vector $\\boldsymbol{\\alpha}$, which we use as the optimization parameter.\n\nRecurrence entanglement purification protocols might include symmetric and asymmetric single-qubit gates before and after the bilateral action of the two-qubit operations \\cite{Bennett1,Deutsch}. These are important in an analytical approach to maintain the form of the state after each iteration, however, entanglement is not affected by this process. In this work, we shall omit these single-qubit gates as we are focused on increasing the value of the concurrence and not on the specific form of the state after each iteration step. \n\n\\subsection{Protocol optimization}\\label{ProtOpt}\n\nIn order to optimize the protocol described in Sec.~\\ref{EntPurif}, we first define an appropriate parametrization for the two-qubit unitary operations used in Eq.~\\eqref{eq: maprho}. In principle, one should consider elements from $U(4)$, the group of $4 \\times 4$ unitary matrices, which contains the subgroup $SU(4)$. However, $U(4)$ is the semi-direct product of $U(1)$ and $SU(4)$, where elements of $U(1)$ are rotations of the unit circle \\cite{Baker}. Therefore,\nchoosing elements from $SU(4)$ in Eq.~\\eqref{eq: maprho}\nrepresents the most general unitary quantum operation involving two-qubit gates at locations $A$ and $B$.\nElements in $SU(4)$ can be parametrized in the following way \\cite{Tilma1}:\n\\begin{eqnarray} \\label{eq: eulerU}\n U(\\boldsymbol{\\alpha})=e^{i \\sigma_3 \\alpha_1} e^{i \\sigma_2 \\alpha_2} e^{i \\sigma_3 \\alpha_3} e^{i \\sigma_5 \\alpha_4} e^{i \\sigma_3 \\alpha_5} e^{i \\sigma_{10} \\alpha_6} \n e^{i \\sigma_3 \\alpha_7} e^{i \\sigma_2 \\alpha_8} \\nonumber \\\\ \n e^{i \\sigma_{3} \\alpha_9} e^{i \\sigma_5 \\alpha_{10}} e^{i \\sigma_3 \\alpha_{11}} e^{i \\sigma_{2} \\alpha_{12}} \n e^{i \\sigma_3 \\alpha_{13}} e^{i \\sigma_8 \\alpha_{14}} e^{i \\sigma_{15} \\alpha_{15}}, \\nonumber \n\\end{eqnarray}\nwith $\\boldsymbol{\\alpha}=(\\alpha_1, \\alpha_2, \\dots, \\alpha_{15})^T \\in \\mathbb{R}^{15}$ ($T$ denotes the transposition), and\n\\begin{eqnarray}\n && 0 \\leqslant \\alpha_1, \\alpha_3, \\alpha_5, \\alpha_7, \\alpha_9, \\alpha_{11}, \\alpha_{13} \\leqslant \\pi, \\nonumber \\\\\n && 0 \\leqslant \\alpha_2, \\alpha_4, \\alpha_6, \\alpha_8, \\alpha_{10}, \\alpha_{12} \\leqslant \\frac{\\pi}{2} \\nonumber \\\\\n && 0 \\leqslant \\alpha_{14} \\leqslant \\frac{\\pi}{\\sqrt{3}}, \\quad 0 \\leqslant \\alpha_{15} \\leqslant \\frac{\\pi}{\\sqrt{6}}, \\label{eq:anglecond}\n\\end{eqnarray}\nwhere $\\sigma_i,\\ i=1, ..., 15$ form a Gell-Mann type basis of the Lie group $SU(4)$, see Appendix \\ref{AppendixA}. This is called the Euler angle parametrization of $SU(4)$, which is sufficient\nto represent every element of the Lie group. For example, the canonical parametrization $e^{i H}$, where $H$ is a $4 \\times 4$ self-adjoint matrix with trace zero, is not minimal, because after exponentiation we may have multiple wrappings around the great circles of the $7$-sphere. \n\nOur aim is to increase the concurrence, which is a non-linear function of the state $\\rho$ and the protocol's unitary matrix $U$. Furthermore, we have lower and upper bounds on $\\boldsymbol{\\alpha}$, as given in Eq.~\\eqref{eq:anglecond}. We then consider a cost function $f:\\mathbb{R}^{15} \\rightarrow \n\\mathbb{R}$ as\n\\begin{align}\n f(\\boldsymbol{\\alpha}) = 1 - \\mathcal{C}'(\\boldsymbol{\\alpha}) \\label{eq:f}\n\\end{align}\nand our optimization problem is, to find local minimizers, e.g., a point $\\boldsymbol{\\alpha}^*$ such that\n\\begin{equation}\n f(\\boldsymbol{\\alpha}^*)\\leqslant f(\\boldsymbol{\\alpha}) \\nonumber\n\\end{equation}\nfor all $\\boldsymbol{\\alpha}$ in the Euclidean norm defined neighborhood of $\\boldsymbol{\\alpha}^*$. The gradient $\\nabla f (\\boldsymbol{\\alpha})$ can be made available to us due to an automatic differentiation \\cite{JAX}, so we \nimplement for the optimization a quasi-Newton algorithm the so-called limited memory Broyden-Fletcher-Goldfarb-Shanno (L-BFGS) method \\cite{Liu}. Here, we have a constrained optimization problem, because $\\boldsymbol{\\alpha}$ takes values in the hyperrectangle defined by Eq.~\\eqref{eq:anglecond} and thus we use the L-BFGS approach of Ref. \\cite{Ryrd}. \nAs long as the state to be purified is given, the above approach yields an optimal $\\boldsymbol{\\alpha}^*$ and a concurrence $\\mathcal{C}'$ as close as possible to one. \nThe vector $\\boldsymbol{\\alpha}^*$ gives the best two-qubit unitary gate associated with this given state. However, our main aim is to provide the best gate, which is capable to purify most effectively certain class of states and not only a fixed one. This is relevant in a scenario where the generation of distant entanglement is affected by noise, leading to slightly different types of states entering the protocol.\nIn order to formulate this quantitatively we introduce a probability density function (PDF) $p(\\boldsymbol{x})$, where the vector $\\boldsymbol{x}$ defines uniquely the state $\\rho$ according to a parametrization. \nFor example, in the case of a Werner state \\cite{Werner}\n\\begin{eqnarray}\n\\hat{\\rho}(x) &=& x \\ket{\\Psi^{-}}\\bra{\\Psi^{-}} + \\frac{1-x}{3} \\ket{\\Psi^{+}}\\bra{\\Psi^{+}} \\nonumber \\\\\n&+& \\frac{1-x}{3} \\ket{\\Phi^{-}}\\bra{\\Phi^{-}}+ \\frac{1-x}{3} \\ket{\\Phi^{+}}\\bra{\\Phi^{+}}, \n\\label{eq:wernerstate}\n\\end{eqnarray} \nwe have $x\\in[0,1]$ and the PDF satisfies\n\\begin{equation}\n \\int^1_0 p(x)=1. \\nonumber\n\\end{equation}\nIn general, a two-qubit state can be described by $15$ parameters, which have to fulfill some non-trivial conditions \\cite{Kimura,Byrd,Gamel}. The choice of the PDF is not straightforward and the only guideline \nwe have is that the support $\\operatorname{supp}(p)$ consists of all $\\boldsymbol{x}$, which define entangled states. This is motivated by \nthe fact that separable states are not purifiable. In the case of the Werner states, the support of the PDF is the interval $(0.5,1]$. Thus, the PDF may put more weight on states with a given concurrence.\n\\begin{algorithm}[H]\n\t\t\\setstretch{1.20}\n\t\t\\caption{Optimization of recurrence protocol}\n\t\t\\label{alg:ouralgo}\n\t\t\\hspace*{\\algorithmicindent} \\textbf{Input} $\\boldsymbol{\\rho}(\\boldsymbol{x})$, optimizer $\\text{OPT}$\\\\\n\t\t\\hspace*{\\algorithmicindent} \\textbf{Output} $\\boldsymbol{\\rho}(\\boldsymbol{x})$\n\t\t\\begin{algorithmic}[1]\n\t\t\\For{$j=1$ to $M$}\n\t\t\\State $\\boldsymbol{x}_j \\sim p(\\boldsymbol{x}_j)$\n\t\t\\State $\\boldsymbol{\\rho}_j = \\boldsymbol{\\rho}(\\boldsymbol{x}_j)$\n\t\t\\EndFor\n\t\t\t\\For{$i=1$ to $N$}\n\t\t\t\\State $U_{AB}(\\boldsymbol{\\alpha}) = U_{A1,A2}(\\boldsymbol{\\alpha}) U_{B1,B2}(\\boldsymbol{\\alpha})$\n\t\t\t\\For{$j=1$ to $M$}\n\t\t\t\\State $\\boldsymbol{\\sigma}^{j}(\\boldsymbol{\\alpha}) = U_{AB}(\\boldsymbol{\\alpha}) \\boldsymbol{\\rho}_j U^{\\dagger}_{AB}(\\boldsymbol{\\alpha})$\n\t\t\t\\For{$l=1$ to $4$}\n\t\t\t\\State $\\boldsymbol{\\sigma}^{jl}_A(\\boldsymbol{\\alpha}) = \\frac{\\bra{l} \\boldsymbol{\\sigma}^{j}(\\boldsymbol{\\alpha}) \\ket{l} }{\\mathrm{Tr}\\{\\ketbra{l}{l} \\boldsymbol{\\sigma}^{j}(\\boldsymbol{\\alpha}) \\}}$\n\t\t\t\\EndFor\n\t\t\t\\EndFor\n\t\t\t\\State $\\mathcal{C}^l(\\boldsymbol{\\alpha}) = \\frac{1}{M}\\sum_{j=1}^M \\mathcal{C}(\\boldsymbol{\\sigma}^{jl}_A)$\n\t\t\t\\State $l' = \\underset{l=1,2,3,4}{\\text{argmin}}\\ \\left(1 - \\mathcal{C}^l(\\boldsymbol{\\alpha})\\right)$\n\t\t\t\\State $\\bar{f}(\\boldsymbol{\\alpha}) = 1 - \\mathcal{C}^{l'}(\\boldsymbol{\\alpha})$\n\t\t\t\\State $\\boldsymbol{\\alpha}^* = \\text{OPT}(\\bar{f}(\\boldsymbol{\\alpha}), \\nabla_{\\boldsymbol{\\alpha}} \\bar{f}(\\boldsymbol{\\alpha}))$\n\t\t\t\\State $\\boldsymbol{\\rho}_j = \\boldsymbol{\\sigma}^{jl'}_{A}(\\boldsymbol{\\alpha}^*) \\otimes \\boldsymbol{\\sigma}^{jl'}_{A}(\\boldsymbol{\\alpha}^*)$\n\t\t\t\\EndFor\n\t\t\\end{algorithmic}\n\t\\end{algorithm}\nIn this context, we have an output concurrence $\\mathcal{C}'(\\boldsymbol{\\alpha}, \\boldsymbol{x})$ depending on both the two-qubit gate and the input state. Therefore, we are going to use an average cost function \n\\begin{equation}\n \\bar{f}(\\boldsymbol{\\alpha}) = 1 - \\int_{\\operatorname{supp}(p)} \\mathcal{C}'(\\boldsymbol{\\alpha}, \\boldsymbol{x}) p(\\boldsymbol{x})\\, d \\boldsymbol{x}, \\label{eq:cost}\n\\end{equation}\nin the L-BFGS algorithm. The integral can be either solved numerically or approximated via a quasi-Monte Carlo approach by sampling $\\boldsymbol{x}$ over its corresponding parameter distribution $p(\\boldsymbol{x})$:\n\\begin{align}\n \\bar{f}(\\boldsymbol{\\alpha}) = 1 - \\underset{\\boldsymbol{x} \\sim p(\\boldsymbol{x})}{\\mathbb{E}} \\left[\\mathcal{C}'(\\boldsymbol{\\alpha}, \\boldsymbol{x}) \\right].\n\\end{align}\nWe chose the second approach since it proved to be precise enough and numerically faster. A summarize of the optimization routine can be seen in Algorithm~\\ref{alg:ouralgo}. \n\nNow, we shed light on the meaning of the average cost function through the following two examples. We employ the CNOT gate\n\\begin{equation}\n U_{\\text{CNOT}}=\\begin{pmatrix} 1 & 0 & 0 & 0 \\\\ 0 & 1 & 0 & 0 \\\\ 0 & 0 & 0 & 1\\\\ 0 & 0 & 1 & 0 \\end{pmatrix}=\n e^{i\\frac{3\\pi}{4}} \\times \\underbrace{U'}_{\\in SU(4)}, \\nonumber\n\\end{equation}\nwhere $U'$ in the Euler angle parametrization is given by setting\n\\begin{eqnarray}\n &&\\alpha_3= \\alpha_5=\\alpha_7= \\frac{\\pi}{4}, \\nonumber \\\\\n &&\\alpha_4= \\alpha_6=\\alpha_{10}= \\frac{\\pi}{2}, \\nonumber\n\\end{eqnarray}\nand the remaining $9$ angles to zero in Eq.~\\eqref{eq: eulerU}. By fixing the vector $\\boldsymbol{\\alpha}$, one is able to get a value for the average cost function of the CNOT gate and thus to evaluate its performance. \\\\\n\n{\\it Example 1.}\nLet us consider the state\n\\begin{equation}\n \\frac{1}{6}\\begin{pmatrix} 1+2x & 0 & 0 & 1-4x \\\\ 0 & 2-2x & 0 & 0 \\\\ 0 & 0 & 2-2x & 0\\\\ 1-4x & 0 & 0 & 1+2x \\end{pmatrix}, \\quad \\text{with} \\quad x \\in [0,1] \\label{eq:example1}\n\\end{equation}\nsubject to the purification protocol with CNOT gates. This state is a rotated Werner state and its concurrence reads\n\\begin{equation}\n \\mathcal{C}(x)=\\begin{cases}2x-1 & x \\in (0.5,1], \\\\ 0 & x \\in [0, 0.5].\\end{cases} \\nonumber\n\\end{equation}\nIn fact, this example is based on the pioneering protocol of Ref. \\cite{Bennett}. \nThe output reads\n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x)=\\frac{3(4x^2-1)}{5-4x+8x^2}, \\quad \\text{for} \\quad x \\in (0.5,1], \\nonumber\n\\end{equation}\nwith success probability\n\\begin{equation}\n P_{\\text{CNOT}}=\\frac{5-4x+8x^2}{9}. \\nonumber\n\\end{equation}\nFor the sake of simplicity, we consider a uniform PDF $p(x)$ with $\\operatorname{supp}(p)=(0.5,1]$. Hence, the input average cost function reads\n\\begin{equation}\n \\bar{f}_{\\text{input}}=1-\\int^1_{0.5} \\mathcal{C}(x) p(x)\\, dx =0.5, \\nonumber\n\\end{equation}\nand the application of the purification protocol with CNOT gates yields\n\\begin{equation}\n \\bar{f}_{\\text{CNOT}}=1-\\int^1_{0.5} \\mathcal{C}'(x) p(x)\\, dx =0.450103. \\nonumber\n\\end{equation}\nThe result shows that the protocol with two identical copies of states allows one to increase on average the entanglement of the output states.\\\\\n\n{\\it Example 2.}\nNow, we consider the state\n\\begin{equation}\n \\begin{pmatrix} \\frac{x}{2} & 0 & 0 & -\\frac{x}{2} \\\\ 0 & 0 & 0 & 0 \\\\ 0 & 0 & 1-x & 0\\\\ -\\frac{x}{2} & 0 & 0 & \\frac{x}{2} \\end{pmatrix}, \\quad \\text{with} \\quad x \\in [0,1]. \\label{eq:example2}\n\\end{equation}\nThis state with concurrence $\\mathcal{C}(x)=x$ is perfectly purifiable in just one iteration of the protocol \\cite{Torres}. Hence, \n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x)=1, \\nonumber\n\\end{equation}\nwith success probability\n\\begin{equation}\n P_{\\text{CNOT}}=\\frac{x^2}{2}. \\nonumber\n\\end{equation}\nIt is immediate from Eq.~\\eqref{eq:cost} that for {\\it any} PDF with $\\operatorname{supp}(p)=[0,1]$\n\\begin{equation}\n \\bar{f}_{\\text{CNOT}}=1-\\int^1_{0} \\mathcal{C}'(x) p(x)\\, dx =0. \\nonumber\n\\end{equation}\nThis means that the CNOT gate is optimal for this family of states.\n\n\n{\\it Example 3.}\nAs a last example, let us consider the initial state\n\\begin{equation}\n x\\ketbra{\\Phi^+}{\\Phi^+}+(1-x)\\ketbra{\\Phi^-}{\\Phi^-}, \\quad \\text{with} \\quad x\\in[0,1], \n\\end{equation}\nand concurrence $\\mathcal{C}(x)=|1-2x|$. After one iteration of the purification protocol with bilaterally applied CNOT gates, it is not hard to realize that the resulting state has the concurrence \n\\begin{equation}\n\\mathcal{C}'_{\\text{CNOT}}(x)=(1-2x)^2\\nonumber \n\\end{equation}\nwith success probability $P_{\\text{CNOT}}=1$. The output\nconcurrence $\\mathcal{C}'_{\\text{CNOT}}(x)$ is less than or equal to $\\mathcal{C}(x)$ for all $x \\in [0,1]$. Thus, we conclude by using the properties of concurrence and integration that\n\\begin{equation}\n \\int^1_{0} \\mathcal{C}'_{\\text{CNOT}}(x) p(x)\\, dx \\leqslant \\int^1_{0} \\mathcal{C}(x) p(x)\\, dx\n\\end{equation}\nfor {\\it any} PDF with $\\operatorname{supp}(p)=[0,1]$. Hence, the initial average cost function $\\bar f_{\\rm input}$ is always less than or equal to \n$\\bar f_{\\rm CNOT}$. This is an example where the purification fails with the sole implementation of the CNOT gate. It must be noted that\nfor $x\\neq 0.5$, the state in this example can be purified with previous protocols \\cite{Bennett1,Deutsch} that work on the Bell basis, and where the implementation with the CNOT gate is now accompanied by local single-qubit gates.\n\nThese examples demonstrate the meaning of the average cost function. It is obvious that a two-qubit gate can be optimal for certain family of states and less optimal for other type of state.\nIn the subsequent section, we will investigate numerically several cases. \n\n\\begin{figure*}[t!]\n \\begin{center}\n \\includegraphics[width=.39\\textwidth]{fig2a.pdf}\n \\includegraphics[width=.39\\textwidth]{fig2b.pdf}\n \\includegraphics[width=.39\\textwidth]{fig2c.pdf}\n \\includegraphics[width=.39\\textwidth]{fig2d.pdf}\n\\caption{Optimized purification protocol for the family of states in Eq.~\\eqref{eq:example1}.\nTop left panel: average cost function $\\bar{f}$ with a uniform PDF as a function of $N$, the number of iterations. Top right panel: The concurrence $\\mathcal{C}$ as a function of $x$. \nFour curves are presented for different values of iterations $N$: $0$ or the input concurrence (dashed line), $1$ (dotted line), $2$ (dash-dotted line), and $3$ (full line).\nBottom left panel: Success probabilities as a function of $x$ for different values of iterations $N$: $1$ (dotted line), $2$ (dash-dotted line), and $3$ (full line). Bottom right panel: The overall success \nprobability after $N=1$ (dotted line), $N=2$ (dash-dotted line), and $N=3$ (full line) iterations as a function of $x$.}\n \\label{fig:TWerner}\n \\end{center}\n\\end{figure*}\n\\section{Results}\n\\label{Results}\n\nIn this section it is demonstrated how our proposed method can find optimal purification schemes. Our first case is the continuation of {\\it Example 1.} in Sec. \\ref{Method}. We have seen so far that \nthe CNOT gate in $N=1$ purification round can reduce the average cost function $\\bar{f}$ approximately by $0.05$. The simulation results with the input state in Eq.~\\eqref{eq:example1} show that the optimal\n$SU(4)$ gate for $N=1$ has a similar improvement on $\\bar{f}$ as the CNOT gate, see Fig.~\\ref{fig:TWerner}. The optimal gates found for $N=2$ and $3$ can further reduce $\\bar{f}$, however, it is easy to check that \nthe CNOT gate is not optimal anymore for these rounds of iterations. In Fig.~\\ref{fig:TWerner} it is also shown that a higher number of iterations result in more concave shapes of the corresponding concurrences. It is \nworth noting that the success probability of the second iteration is lower than the success probability of the first iteration. The overall success probability of $3$ iterations, \nshown also in Fig.~\\ref{fig:TWerner}, is defined as: in the first iteration $4$ qubit pairs, in the second iteration $2$ qubit pairs, and in the third iteration the final qubit pair are successfully purified. These \nresults show that states close to maximally entangled states can be produced already in the third iteration, but the overall success probability of the procedure is getting smaller with the number of \niterations. The only fixed point is $\\mathcal{C}=1$. Thus our method seems to provide the same results to the one found in \\cite{Bennett}, which is based completely on the CNOT gate. As a result, we have investigated the circumstances where the CNOT gate is also optimal. It turns out that the local unitary transformation introduced by \\cite{Deutsch} and given by\n\\begin{equation}\n b_{A_1}^{\\dagger}\\otimes b_{A_2}^{\\dagger}\\otimes b_{B_1}\\otimes b_{B_2}, \\label{eq:localtr}\n\\end{equation}\nwhere\n\\begin{equation}\n b=\\frac{\\mathbb{I}_2+i\\sigma_x}{\\sqrt2} \\nonumber\n\\end{equation}\nwith the Pauli matrix $\\sigma_x$ and the identity map $\\mathbb{I}_2$, is crucial for the CNOT gate. This is the transformation, which transforms a Werner state in Eq.~\\eqref{eq:wernerstate} into \nthe state in Eq.~\\eqref{eq:example1}. Now, if one applies Eq.~\\eqref{eq:localtr} before all iterations of the protocol involving only the CNOT gate, then the same curves are obtained as in Fig.~\\ref{fig:TWerner}. \nThis means that there are more optimal protocols, which yield the same results, and our approach can find them. Thus, the analytically tractable CNOT-based protocol can be used as a benchmark to test \nthe numerical accuracy of our method. Our numerics show at least a $2$ decimal accuracy until the third iteration, but up to the fourth iteration, there is a breakdown in the numerical accuracy. \n\nThe one-parameter family of states in Eq.~\\eqref{eq:example1} has the same concurrence as the Werner state. Therefore, we consider the Werner state to be our next application. In Fig.~\\ref{fig:Werner} \nnumerical results are presented for the average cost function $\\bar{f}$ and the overall success probability which exhibit the same behavior found for the one-parameter family of states in Eq.~\\eqref{eq:example1}. \nIn this case, one can note an improvement of $0.04$ in $\\bar{f}$ after $N=3$ iterations. This is less than the previously obtained value of $0.09$ shown in Fig.~\\ref{fig:TWerner}.\nFurthermore, this is accompanied by another numerical \ninaccuracy: the overall success probability at $\\mathcal{C}=1$ is less than one. Given these results, the proposed method can find optimal entangling two-qubit gates for at least two iterations. It is also clear \nfrom these tests that the algorithm is always reducing $\\bar{f}$, but from $N=3$ iterations this might be not an optimal improvement of the concurrence. This originates from the fact that the gradient $\\nabla f (\\boldsymbol{\\alpha})$, see Eq.~\\eqref{eq:f}, is almost flat in the neighborhood of $\\mathcal{C}=1$ for $N>2$ iterations and thus the numerical search for the optimal gate, i.e, \nthe search for $\\boldsymbol{\\alpha}^* \\in \\mathbb{R}^{15}$, becomes inefficient. \n\n\\begin{figure*}[t!]\n \\begin{center}\n \\includegraphics[width=.39\\textwidth]{fig3a.pdf}\n \\includegraphics[width=.39\\textwidth]{fig3b.pdf}\n\\caption{Optimized purification protocol for Werner states, see Eq.~\\eqref{eq:wernerstate}.\nLeft panel: average cost function $\\bar{f}$ with a uniform PDF as a function of $N$, the number of iterations. Right panel: The overall success \nprobability after $N=1$ (dotted line), $N=2$ (dash-dotted line), and $N=3$ (full line) iterations as a function of $x$.}\n \\label{fig:Werner}\n \\end{center}\n\\end{figure*}\n\nInspired by {\\it Example 2.}, we consider the state in Eq.~\\eqref{eq:example2} and also\n\\begin{equation}\nx\\ketbra{\\Psi^-}{\\Psi^-}+(1-x)\\ketbra{\\Upsilon}{\\Upsilon} \\label{eq:MaZs}\n\\end{equation}\nwith $x \\in [0,1]$ and \n\\begin{equation}\n \\ket{\\Upsilon}=\\frac{1}{\\sqrt2} \\left(\\ket{\\Psi^+}+i\\ket{\\Phi^-}\\right). \\nonumber\n\\end{equation}\nBy using the unitary transformation in Eq.~\\eqref{eq:localtr} on Eq.~\\eqref{eq:MaZs} one obtains Eq.~\\eqref{eq:example2} and therefore it is expected that the only optimal two-qubit gates are the ones that purify these states\nin one iteration. Both states have an input concurrence $\\mathcal{C}(x)=x$. However, it is important to note that the CNOT gate is not optimal for the state in Eq.~\\eqref{eq:MaZs}, because after one iteration round\n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x)=\\begin{cases} 0& x \\in [0,0.5], \\\\ 2x \\frac{2x-1}{1+x^2} & x \\in (0.5,1], \\end{cases} \\nonumber\n\\end{equation} \nand $\\mathcal{C}'_{\\text{CNOT}}(x) \\leqslant x$. The success probability is\n\\begin{equation}\n P_{\\text{CNOT}}=\\frac{1+x^2}{2}. \\nonumber\n\\end{equation}\nThus, the CNOT gate based purification protocol impairs the concurrence. In contrast to this analytical observation, numerical analysis with uniform PDF yields already in the first iteration for both states an average cost function $\\bar{f}\\approx 0.0002$. To demonstrate the robustness of the numerical approach we provide examples of three non-uniform PDFs for $N=1$ \niteration. First, we consider\n\\begin{equation}\np(x)=2x,\\quad \\text{with} \\quad x \\in [0,1], \\nonumber\n\\end{equation}\nwhich describes a situation, where states with higher concurrences are more likely to be subject to the purification. The resulting average cost function is $\\bar{f}\\approx 0.000004$. Secondly, we take \n\\begin{equation}\np(x)=2(1-x),\\quad \\text{with} \\quad x \\in [0,1], \\nonumber\n\\end{equation}\nwhich puts more weight on states with low concurrences and find $\\bar{f}\\approx 0.0005$. Finally, we investigate a PDF\n\\begin{equation}\np(x)=6x(1-x),\\quad \\text{with} \\quad x \\in [0,1], \\nonumber\n\\end{equation}\ni.e., the states around the concurrence $\\mathcal{C}(x)=0.5$ are more likely to participate in the purification, and obtain $\\bar{f}\\approx 0.00005$. These results demonstrate the effectiveness of our approach and up to a numerical precision these one-parameter families of states can be purified in one iteration.\n\n\\begin{figure*}[t!]\n \\begin{center}\n \\includegraphics[width=.39\\textwidth]{fig4a.pdf}\n \\includegraphics[width=.39\\textwidth]{fig4b.pdf}\n\\caption{Optimized purification protocol for the state in Eq.~ \\eqref{eq:qrstate}. Left panel: Average cost function $\\bar{f}$ with a uniform PDF as a function of $N$, the number of iterations. \nCrosses are the results of the CNOT gate, whereas dots display the numerical optimization. They have been connected by lines to guide the eye. Right panel: The concurrence $\\mathcal{C}'$\nas a function of $x$ and $y$ after one purification round. The cone-type surface is obtained with our approach. A less optimal surface with minima at $x=0$ and along the $y$ axis is the result of the purification protocol \nwith the CNOT gate.}\n \\label{fig:XY}\n \\end{center}\n\\end{figure*}\n\nNext, we consider the following two-parameter family of states \\cite{Bernad2}\n\\begin{equation}\n\\hat{\\rho}(x,y) = \\frac{1}{4}\n\\left(\n\\begin{array}{cccc}\n 1-x & i y & -i y & x-1 \\\\\n -i y & x+1 & -x-1 & i y \\\\\n i y & -x-1 & x+1 & -i y \\\\\n x-1 & -i y & i y & 1-x \\\\\n\\end{array}\n\\right) \\label{eq:qrstate}\n\\end{equation}\nwith $x,y \\in \\mathbb{R}$ and $x^2+y^2\\leqslant1$. The concurrence of this state is $\\sqrt{x^2+y^2}$. In order to relate the performance of our approach, we apply the CNOT gate based purification \nprotocol to this family of states. One obtains\n\\begin{eqnarray}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)&=&\\frac{2|x|}{1+x^2}, \\label{eq:xyCNOTc} \\\\\n P_{\\text{CNOT}}&=&\\frac{1+x^2}{2}, \\label{eq:xyCNOTp}\n\\end{eqnarray}\nafter $N=1$ and\n\\begin{eqnarray}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)&=&\\frac{4 |x| (1+x^2)}{1+6x^2+x^4}, \\label{eq:xyCNOTc2} \\\\\n P_{\\text{CNOT}}&=&\\frac{1+6x^2+x^4}{2 (1+x^2)^2}, \\label{eq:xyCNOTp2}\n\\end{eqnarray}\nafter $N=2$ purification rounds, respectively. If one applies the unitary transformation in Eq.~\\eqref{eq:localtr} on the state before the bilateral CNOT gates are performed, then the above results remain \nunchanged. These results demonstrate that the CNOT gate and the additional tricks, which yielded optimal purifications for the one-parameter family of states, are not optimal in this scenario, since they are outperformed on average by our optimized protocol. \nWe assume that CNOT-based purification protocols cannot exploit all the useful entanglement, because the concurrence \nand the success probability become independent of the $y$ variable. Using these analytical results and the uniform PDF\n\\begin{equation}\n p(x,y)=\\frac{1}{\\pi},\\quad \\text{with} \\quad x^2+y^2 \\leqslant 1, \\nonumber\n\\end{equation}\nwe compare our approach with the above-presented analytical results, see Eq.~\\eqref{eq:xyCNOTc} and Eq.~\\eqref{eq:xyCNOTc2}. In Fig.~\\ref{fig:XY}, we show that our optimized protocol provides after one iteration a $x$ and $y$-dependent concurrence. Although the \nnumerically-obtained concurrence is larger at $x=0$ and $y \\in [-1,1]$ than the surface given by Eq.~\\eqref{eq:xyCNOTc}, one can also observe the opposite at $y=0$ and $x \\in [-1,1]$. However, the concurrence presents a better improvement with\nour algorithm and the average cost function with the uniform PDF stays for more iterations lower than the CNOT-based protocol, see the left figure in panel \\ref{fig:XY}. \nIn regard to the overall success probability, we let our algorithm work until the third iteration and in Fig.~\\ref{fig:XYos} the results are compared with the CNOT-based purification protocol. The obtained \nsurfaces differ only by a $\\pi\/2$ rotation around the $z$-axis. Therefore, there is not much difference in the overall success probability and from this aspect the CNOT gate can also be considered optimal, though, \nit still can not improve the concurrence as effectively as the gate obtained with our approach. However, one might think with a proper choice of local unitary transformations, like the transformation \nin Eq.~\\eqref{eq:localtr} used for Werner and one-step purifiable states, the application of the CNOT gate may result in an optimal purification protocol. Let us consider then two general local unitary \ntransformations at locations $A$ and $B$ for the preparation of two-qubit states. It is enough to consider unitary transformations in $SU(2)$, for which the Euler angle parameterization reads \\cite{Tilma2}\n\\begin{eqnarray}\n U_A&=& e^{i \\sigma_z \\alpha_1} e^{i \\sigma_y \\alpha_2} e^{i \\sigma_z \\alpha_3}, \\\\\n U_B&=& e^{i \\sigma_z \\beta_1} e^{i \\sigma_y \\beta_2} e^{i \\sigma_z \\beta_3},\n\\end{eqnarray}\nwhere $\\sigma_y$ and $\\sigma_z$ are the Pauli matrices. Here, $\\alpha_1, \\beta_1 \\in [0, \\pi]$, $\\alpha_2,\\beta_2 \\in [0, \\pi\/2]$, and $\\alpha_3, \\beta_3 \\in [0, 2\\pi]$ are the Euler angles for $SU(2)$. For the first\niteration, we analyze the output\nunitary of the optimized protocol with \\textit{Mathematica} \\cite{CAS} and observe that the angles $\\alpha_1$ and $\\beta_1$ do not affect the efficiency of the CNOT-based protocol. Optimal \nconcurrences for the remaining four angles yield either the result in Eq.~\\eqref{eq:xyCNOTc} or\n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)=\\frac{2|y|}{1+y^2}, \n\\end{equation}\nwith success probability\n\\begin{equation} \n P_{\\text{CNOT}}=\\frac{1+y^2}{2}, \n\\end{equation}\nfor $\\alpha_2=\\frac{\\pi}{8}$, $\\beta_2=\\frac{3 \\pi}{8}$, and $\\alpha_3=\\beta_3=\\frac{3\\pi}{4}$. For these results the average cost function with uniform PDF yields $\\bar{f}\\approx 0.372$, which is still lower than\nthe average concurrence found by our approach. These results demonstrate that even with local unitary transformations the CNOT gate is globally not optimal for all values of $x$ and $y$. However, when $x$ and $y$\nare known beforehand, then one can use the following local unitary transformation\n\\begin{eqnarray}\n U^\\dagger_A \\otimes U_B, \\quad \\text{with} \\quad U=\\cos(\\theta)\\mathbb{I}_2 + \\sin(\\theta) \\sigma_x, \\label{eq:statedeptr}\n\\end{eqnarray}\nwhere the angle $\\theta$ is a function of $x$ and $y$. For example, for $x,y \\geqslant 0$, we have\n\\begin{equation}\n \\cos(\\theta)=\\sqrt{\\frac{1}{2}+\\sqrt{\\frac{x+\\sqrt{x^2+y^2}}{8 \\sqrt{x^2+y^2}}}}. \\nonumber\n\\end{equation}\nUsing this transformation before the purification protocol, one obtains the state\n\\begin{equation}\n\\frac{1+ \\sqrt{x^2+y^2}}{2} \\ketbra{\\Psi^-}{\\Psi^-}+\\frac{1- \\sqrt{x^2+y^2}}{2} \\ketbra{\\Phi^-}{\\Phi^-}. \\label{eq:Spec}\n\\end{equation}\nNow, the CNOT-based purification yields\n\\begin{equation}\n \\mathcal{C}'_{\\text{CNOT}}(x,y)=\\frac{2\\sqrt{x^2+y^2}}{1+x^2+y^2}, \n\\end{equation}\nwith success probability\n\\begin{equation}\n P_{\\text{CNOT}}=\\frac{1+x^2+y^2}{2}. \n\\end{equation}\nIt is immediate that $\\mathcal{C}'_{\\text{CNOT}}(x,y) \\geqslant \\sqrt{x^2+y^2}$, i.e., we are improving the concurrence. In this particular case, one does not need the average cost function in the numerical search, \nbecause the state is fixed. Our algorithm running with a single state with parameters $(x,y)$ as defined in Eq.~\\eqref{eq:qrstate} instead of an ensemble of states can improve the concurrence, but is unable to find the optimal solution presented above, because the transformation\nEq.~\\eqref{eq:statedeptr} together with CNOT gates is a non-symmetrical transformation at nodes $A$ and $B$.\n\nFinally, let us point out that our approach does not take into account operational or memory errors. These always depend on the implementation and if an experiment can provide us models for the errors then \nour approach can be easily extended. Another important experimental input is the PDF $p(\\boldsymbol{x})$, which is usually subject to the method of generating entangled states between locations A and B. Furthermore, \nthis PDF is assigned to the process of choosing a value $\\boldsymbol{x}$ as a random event, i.e, we have the same two two-qubit pairs parameterized by $\\boldsymbol{x}$ before the purification protocol. In effect, \nwe integrate the parameter dependence of the states away and thus our approach yields an optimal two-qubit gate on average. If the value of $\\boldsymbol{x}$ is fixed in an experimental design, then our method \nprovides an optimal two-qubit gate designed for this particular state. The optimal two-qubit gates can be can be realized by three\nCNOT gates and additional single-qubit rotations \\cite{Vidal,Vatan}, and therefore the experimental generation of this gate is possible with high fidelity \\cite{Debnath}. Also in the context of trapped ions or superconducting quantum circuits, the generation of two-qubit entangling gates can be achieved with high precision using the M\\o{}lmer-S\\o{}rensen gate \\cite{MS} or the $\\sqrt{i \\text{SWAP}}$ \\cite{Fan} gate. Since two-qubit unitaries have a straightforward implementation in many physical settings, thanks to quantum compilation \\cite{Martinez}, we argue that quantum computing platforms implementing quantum communication could actually benefit from globally optimal gates for entanglement purification.\n\n\\begin{figure*}[t!]\n \\begin{center}\n \\includegraphics[width=.39\\textwidth]{fig5.pdf}\n\\caption{The overall success probability of the state in Eq.~ \\eqref{eq:qrstate} after and $N=3$ iterations as a function of $x$ and $y$. Both surfaces are similar in appearance. The one having maxima at $x=1$ \nbelongs to the CNOT-based protocol and the other one having maxima at $y=1$ is obtained via the optimized numerical protocol.}\n \\label{fig:XYos}\n \\end{center}\n\\end{figure*}\n\n\\section{Conclusions}\n\\label{Conclusions}\n\nIn the context of entanglement purification and recurrence protocols, we have presented a method to obtain optimal protocols. This method searches for the optimal two-qubit gate, which is applied bilaterally\nat the nodes $A$ and $B$ in order to distill from an ensemble of noisy pairs a higher fidelity state with respect to a maximally entangled state. Our strategy considers situations, where the \nexperimental apparatus does not have enough control over the entanglement generation phase and different experimental runs result in different states. Here, we assume that the same copies of the noisy state can be \ngenerated before the purification protocol takes place, but a different experimental run could result in different states. Errors originating from local operations, memory requirements, or even \nclassical communication are neglected for now.\n\nWe have numerically demonstrated the optimality of our proposal for several states. In the case of the Werner state, we have found several optimal two-qubit gates and their performances are the same as the CNOT-based Bennett protocol \\cite{Bennett}. Thus, for Werner states the optimality cannot be improved beyond the already known performance. Next, we have investigated a family of states that one can purify \nin one step, i.e., two noisy copies of states are enough to obtain a maximally entangled state, where we immediately obtain minimal average cost function $\\bar{f} \\approx 0$, as expected. Finally, we consider a two-parameter family of states which is considered in theoretical models of a quantum repeater \\cite{Bernad2}. Our numerical investigations demonstrate that a\nsingle bilaterally applied CNOT gate cannot be globally optimal for these states. On the other hand, when the state is known, then parameter-dependent, non-symmetric local transformations allow again the CNOT to be also \none of the optimal two-qubit gates. This case exemplifies the difference between protocols that are optimal for a full class of parameterized states and those that are only optimal for a single state. \n\nIn conclusion, we have proposed a general method to optimize entanglement purification for an arbitrary family of states. Our algorithm \\cite{Francesco} can find the most optimal two-qubit gate that on average induces the highest increase of entanglement. We remark that the method is general for the set of parameters defining the states. Optimizing for a single state is also possible, as shown in one of the \npresented examples. Furthermore, we have focused on the degree of entanglement measured by the concurrence and not on the fidelity with respect to a particular Bell state. \nThis is motivated by the fact that there exist an infinite number of maximally entangled states, and therefore entangled mixed states can present a larger fidelity with respect to maximally entangled states that might not correspond to Bell states defined in the computational basis. \nA possible drawback is that one cannot know with certainty the final maximally entangled state produced by the protocol. This and other issues such as different input states, asymmetric two-qubit gates, \nand non-ideal local operations remain open questions. However, this work aims to introduce the concept of globally optimized recurrence protocols that is flexible enough to incorporate the \naforementioned points in further investigations, which we plan to carry out in the future. \n\n\\begin{acknowledgments}\nThis work is supported by the DFG under Germany's Excellence Strategy-Cluster of Excellence Matter and Light for Quantum\nComputing (ML4Q) EXC 2004\/1-390534769. The results were obtained using the libraries JAX \\cite{JAX}, Scipy \\cite{Scipy} and QuTiP \\cite{Qutip}.\n\n\\end{acknowledgments}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Derivation of the path integral}\n\\label{app:pathintegral}\nFor simplicity, in this section, we assume a quantum hamiltonian given by\n\\begin{equation}\n\\hat{H}=\\sul{\\alpha,\\beta}{}h_{\\alpha\\beta}\\hat{c}_{\\alpha}^{\\dagger}\\hat{c}_{\\beta}^{}+\\sul{\\substack{\\alpha,\\beta \\\\ \\alpha\\neq\\beta}}{}U_{\\alpha\\beta}\\hat{c}_{\\alpha}^{\\dagger}\\hat{c}_{\\beta}^{\\dagger}\\hat{c}_{\\beta}^{}\\hat{c}_{\\alpha}^{}.\n\\label{eq:quantum_hamiltonian_app}\n\\end{equation}\nThe result for a non-diagonal interaction $U_{\\alpha\\beta\\gamma\\nu}$, however, is given in appendix \\ref{app:classical_hamiltonians}\nIn order to get from eq.~(\\ref{eq:propagator_in_grassmann_path-integral}) to the complex path integral eq.~(\\ref{eq:propagator_in_complex_path-integral}), the following two integrals with $j,j^\\prime\\in\\mathbb{N}_0$, will be inserted:\n\\begin{align}\n&\\intg{0}{2\\pi}{\\theta}\\intg{}{}{^2\\phi}\\exp\\rbr{-\\abs{\\phi}^2+\\conjg{\\phi}{\\rm e}^{{\\rm i}\\theta}-{\\rm i} j\\theta}\\phi^{j^\\prime}=2\\pi^2\\delta_{j,j^\\prime}\n\\label{eq:inserted_integral1} \\\\\n&\\intg{}{}{^2\\phi}\\intg{}{}{^2\\mu}\\exp\\rbr{-\\abs{\\phi}^2-\\abs{\\mu}^2+\\conjg{\\phi}\\mu}\\phi^j\\rbr{\\conjg{\\mu}}^{j^\\prime}=\\pi^2j!\\delta_{j,j^\\prime},\n\\label{eq:inserted_integral2}\n\\end{align}\nThereby ${\\rm d}^2\\mu={\\rm d}\\Re{\\mu}{\\rm d}\\Im{\\mu}$, \\textit{i.e.}~the integrations over $\\phi$ and, in the second case, over $\\mu$ run over the whole complex plane. One should notice, that the first of these two integrals is just the second one, but with the modulus of $\\mu$ already integrated out.\n\nThe first of these two integrals is used to decouple $\\zetakv{0}$ from $\\zetakv{1}$ by the following identity:\n\n\\begin{widetext}\n\\begin{align}\n\\int&{\\rm d}^{2J}\\zetakj{0}{}\\exp\\rbr{-\\conjg{\\zetakv{0}}\\cdot\\zetakv{0}}\\cbr{\\prodl{j=1}{J}\\rbr{1+\\conjg{\\chikj{0}{j}}\\zetakj{0}{j}}}\\prodl{j=1}{J}\\rbr{\\conjg{\\zetakj{0}{j}}}^{\\nij{j}}= \\nonumber \\\\\n&\\int\\frac{{\\rm d}^{2N_i}\\phikj{0}{}}{\\pi^{N_i}}\\int\\limits_{0}^{2\\pi}\\frac{{\\rm d}^{N_i}\\thetakj{i}{}}{\\left(2\\pi\\right)^{N_i}}\\int{\\rm d}^{2J}\\zetakj{0}{}\\exp\\rbr{-\\conjg{\\zetakv{0}}\\cdot\\zetakv{0}-\\abs{\\phikv{0}}^2+\\conjg{\\phikv{0}}\\cdot\\mukv{0}}\\cbr{\\prodl{j=1}{J}\\rbr{1+\\conjg{\\chikj{0}{j}}\\phikj{0}{j}}}\\cbr{\\prodl{j=0}{J-1}\\rbr{1+\\zetakj{0}{J-j}\\conjg{\\mukj{0}{J-j}}}}\\prodl{j=1}{J}\\rbr{\\conjg{\\zetakj{0}{j}}}^{\\nij{j}},\n\\label{eq:insert_integrals_0}\n\\end{align}\n\\end{widetext}\nwith $\\mukj{0}{j}=\\nij{j}\\exp\\rbr{{\\rm i}\\thetakj{i}{j}}$ for all $j\\in\\{1,\\ldots,J\\}$, where $J$ is the number of single particle states taken into account. Note that here, for the initially unoccupied single particle states, the phases $\\thetakj{i}{j}$ are arbitrary but fixed, \\textit{e.g.}~to zero, while the integration runs only over those initial phases $\\thetakj{i}{j}$, for which $\\nij{l}=1$. In this way, the integrals, that have to be performed exactly, in order to get a reasonable and correct semiclassical approximation for the propagator are already done, and do not have to be carried out later.\n\nFor the $N_i=\\sum_{j=1}^{J}\\nij{j}$ initially occupied single particle states, the identity follows directly from eq.~(\\ref{eq:inserted_integral1}), while for the unoccupied ones, it is important to notice, that the term $\\conjg{\\chikj{0}{j}}\\zetakj{0}{j}$ does vanish when integrating over $\\zetakv{0}$. This is because of the properties of the Grassmann integrals eq.~(\\ref{eq:vanishing_grassmann-integrals}) and the fact, that there is no $\\conjg{\\zetakj{0}{j}}$ for those components, for which $\\nij{j}=0$.\n\nThe thus obtained expression is the starting point for an iterative insertion of integrals of the form of eq.~(\\ref{eq:inserted_integral2}). For $1\\leqm