{"text":"\\section{Introduction}\n\n\nDuring the last two years a race of industrial and research organizations has been opened to develop a ready-to-implement engineering solution for quantum computing (QC). It resulted in the QC market closely resembling the ascent ages of classical computing industry.\nNamely, there were many underdeveloped computing architectures which being incompatible with each other required significant efforts in porting software and algorithmic solutions between them. Given a broadly supported opinion that in the near term we are unlikely to become witnesses to flexible large-scale quantum architectures, there is a critical need to develop portable, architecture-agnostic hybrid quantum-classical frameworks that will allow solving large-scale computational problems on small-scale quantum architectures. \n\n\nThere are multiple emerging quantum computation paradigms. The performance comparison of these paradigms is an important research topic. In this paper, we present for the first time a performance comparison of two leading quantum computation paradigms - D-Wave quantum annealing and gate-based universal quantum computation. Both approaches have great potential for achieving quantum speedup for a number of important problems~\\cite{king2018observation,romero2018strategies,ambainis2018quantum,dunjko2018computational}.\n\n\n\nThe first approach, quantum annealing (QA), is based on adiabatic quantum computation (AQC)~\\cite{mcgeoch2014adiabatic}. QA solves computational problems by using a guided quantum evolution~\\cite{yang2017optimizing}. \nThe evolution starts with an initial Hamiltonian with an easy-to-prepare ground state and ends up in the ground state of the problem Hamiltonian. QA is based on the adiabatic theorem that guarantees that if the Hamiltonian is evolved slowly then transitions to excited states are suppressed during the adiabatic evolution~\\cite{yang2017optimizing}.\nThe D-Wave quantum annealer uses superconducting flux qubits \\cite{amin2004,dwave2018} and has been shown to solve optimization problems on graphs \\cite{ushijima2017graph}, machine learning \\cite{omalley2017}, traffic flow optimization \\cite{neukart2017}, and simulation problems \\cite{harris2018}.\nQuantum and hybrid quantum-classical approaches have been employed.\n\nThe second approach is often referred to as the gate-based or universal QC. This mode of QC was theoretically demonstrated to have a great potential for exponential speedups over best known classical algorithms~\\cite{nielsen2002quantum}. In the near term, the capability of the quantum devices is limited by the number of qubits, low fidelity of gates, and lack of error correction. These limitations constrain us to using low-depth quantum circuits (i.e., quantum circuits with few gates) on a small number of qubits. Within the constraints of near-term intermediate-scale quantum (NISQ) technology~\\cite{preskill2018quantum}, a number of hybrid quantum-classical algorithms were developed and experimentally demonstrated to solve small problems. One of the most promising of such algorithms is Quantum Approximate Optimization Algorithm (QAOA)~\\cite{farhi2014quantum,farhi2016quantum}. QAOA is inspired by adiabatic quantum computation. Similarly to AQC and QA, the evolution path starts with an easy-to-prepare Hamiltonian in the ground state and evolves to the final Hamiltonian that encodes the solution of the problem remaining in the ground state. However, unlike QA in QAOA the evolution is performed by applying a series of parametrized gates called ansatz~\\cite{mcclean2016theory} which is parametrized by a set of variational parameters. This is accomplished by a hybrid approach that combines quantum evolution and classical variational optimization for optimal QAOA parameters~\\cite{yang2017optimizing} with the goal of finding the evolution path that prepares the ground state of the problem Hamiltonian. \n\n\\section{Methodology}\n\nThis work addresses three main challenges. First, we show how to use quantum computing to solve the community detection problem, a well known NP-hard problem. Second, we present an approach to solving realistic large problems using the NISQ hardware with a limited number of noisy qubits. Third, we demonstrate a method that is portable across two leading quantum computation paradigms and can be easily extended to future hardware.\n\nThe community detection problem (or modularity graph clustering) has a variety of applications ranging from biology to social network analysis~\\cite{palla2005uncovering,su2010glay,bardella2016hierarchical,nicolini2017community}. \nIts complexity \\cite{brandes2006maximizing} and practical importance justify an attempt to solve it using QC. The goal of the community detection is to split nodes of a graph $G=(V,E)$ into communities by maximizing its modularity~\\cite{2006PNAS..103.8577N}:\n\n\\begin{equation}\nH = \\frac{1}{4|E|}\\Sigma_{ij}(A_{ij} - \\frac{k_ik_j}{2|E|})s_is_j = \\frac{1}{4|E|}\\Sigma_{ij}B_{ij}s_is_j,\n\\label{eq:mod}\n\\end{equation}\n\n\\noindent where $s_i\\in\\{-1,+1\\}$ are variables indicating node $i$th community assignment, $k_i$ is a degree $i\\in V$, and $A$ is the adjacency matrix of $G$.\nIn this paper, we will focus on clustering the graph into two communities. There are several approaches to generalize the problem for cases when the number of communities is greater than 2.\n\nThe clustering of large networks is currently impossible with existing quantum computers because of the small number of available qubits.\nThis limitation applies both to quantum annealing~\\cite{ushijima2017graph} and universal quantum computing~\\cite{otterbach2017unsupervised}. To tackle large problems using available quantum hardware, we use a hybrid quantum-classical local-search approach. Our approach is inspired by existing numerous local-search heuristics (see~\\cite{rotta2011multilevel} for a review). Our algorithm finds a solution to the global community detection problem by selecting subproblems small enough to fit on the target quantum computer, solving them using a quantum algorithm and iterating until the solution to the global problem is found. The outline is presented in Algorithm~\\ref{alg:outline}.\n\n\\begin{algorithm}\n \\caption{Community Detection}\\label{alg:outline}\n\\begin{algorithmic}\n\\Procedure{Community detection}{Graph $G$}\n\\State solution = initial\\_guess($G$)\n \\While{not converged}\n \t\\State $X$ = populate\\_subset($G$)\n \t\\State \/\/ \\textit{using QAOA or D-Wave QA}\n \t\\State candidate = solve\\_subproblem($G$, $X$)\n \t\\If{$\\mbox{candidate} > \\mbox{solution}$}\n \t\\State solution = candidate\n \\EndIf\n \\EndWhile\n \\EndProcedure\n\\end{algorithmic}\n\\end{algorithm}\n\nIn particular, we start with a random community assignment. At each step we select a subproblem (subset of vertices $X\\subset V$) by taking the vertices with highest potential gain if moving them from one community to another. The gain for each vertex can be computed efficiently~\\cite{2006PNAS..103.8577N}. Then we fix the community assignment of all $i\\not\\in X$, encode them into the problem as boundary conditions (denoted by $\\tilde{s}_j$, a typical technique in many heuristics \\cite{leyffer2013fast,hager2018multilevel}) and maximize\n\n\\begin{equation}\n\\label{eq:subproblem}\n\\arraycolsep=1.4p\n\\begin{array}{r c l}\nQ_{s} & = & \\sum_{i>j | i,j\\in X}2B_{ij}s_is_j + \\sum_{i\\in X}\\sum_{j\\not\\in X}2B_{ij}s_i\\tilde{s}_j \\\\\n & = & \\sum_{i>j| i,j\\in X}2B_{ij}s_is_j + \\sum_{i\\in X}C_{i}s_i. \n\\end{array}\n\\end{equation}\n\nThe subproblems are solved using QC. To satisfy the constraints of available hardware, we fix the subproblem size to some small number (in our experiments, it was 25). \n\n\n\\section{Implementation details and Results}\n\nWe implement our local search algorithm in Python using the graph methods provided by NetworkX~\\cite{hagberg2008}. The novelty of our approach is that it allows to use D-Wave QA, QAOA and classical Gurobi \\cite{optimization2014inc} solvers interchangeably simply by passing different flags, enabling rapid prototyping and direct comparison of different methods as the hardware and its capabilities evolve. Additionally, Gurobi was used as a global optimization solver for the sake of quality comparison. To our knowledge this is the first attempt to directly compare universal quantum computing and quantum annealing. Our framework is also easily extendable, making it possible for researchers to add new backends as they become available. We plan to release the framework as an open-source project.\n\nOur results are presented in Figure \\ref{fig:results}. In these experiments, we used the Intel-QS~\\cite{smelyanskiy2016qhipster} simulator for QAOA (at the time our group did not have access to a universal quantum computer of sufficient size). We use six real-world networks from the KONECT dataset~\\cite{kunegis2013konect} with up to 400 nodes as our benchmark. For each network, we ran 30 experiments with different random seeds. The same set of seeds is used between three backend solvers, making the results directly comparable. The subproblem size is fixed at 25 (i.e., 25 qubits are used). Our results demonstrate that the quantum local search approach with both quantum methods is capable of achieving results comparable to state-of-the-art, with a potential to outperform as hardware evolves.\n\n\n \n\n\n\n\n \\vspace{-0.78cm}\n \n\\begin{figure}[htb]\n \\begin{tikzpicture} \n \\node (img) {\\includegraphics[width=0.8\\linewidth]{.\/fig\/mod_iter.pdf}};\n \\node (temp) [left=of img]{};\n \\node[below=of img, node distance=0cm, yshift=1.1cm,font=\\color{black}] {Network Name};\n \\node[below=of temp, node distance=2cm, anchor=center,yshift=-1.0cm,,xshift=1.2cm, rotate=90,font=\\color{black}] {Num. Solver Calls};\n \\node[above=of temp, node distance=2cm, xshift=1.2cm,yshift=0.8cm, rotate=90, anchor=center,font=\\color{black}] {Modularity};\n \\end{tikzpicture}\n \\vspace{-0.4cm}\n \\caption{Box-plots comparing modularity scores (greater is better) and number of solver calls (less is better) respectively for the three different solvers. For the graph {\\tt oz}, Gurobi and D-Wave returned a modularity score greater than the Global Solver (best known value)}\n \\label{fig:results}\n\\end{figure}\n \\vspace{-0.6cm}\n\\section{Discussion}\n\n\n\n\nIn the near term, quantum hardware will be in a constant state of change. Many different NISQ-era hardware solutions will appear and some will be abandoned. In the midst of such evolutionary times, we want to be able to continue research in quantum algorithms and head towards solving real-world problems. To accomplish this, we need portable, architecture-agnostic hybrid quantum-classical frameworks that will allow solving large-scale computational problems on small-scale quantum architectures. Moreover, these frameworks need to be robust and future-proof. In this work, we have presented a prototype of such a framework for solving the problem of community detection in networks on two distinctively different architectures: D-Wave quantum annealer and universal quantum computer. We suggest extending this approach for solving other types of problems in science.\n\nThe constant change of hardware and overall immaturity of the existing technology leads to many risks in QC. In spite of major effort, it has not been experimentally demonstrated yet an ability to achieve speedups over state-of-the-art classical supercomputers and there are valid concerns about scalability of existing implementations~\\cite{brugger2018quantum}. Advances in material design and engineering will allow the community to overcome those hurdles. We expect QC to eventually become a part of the HPC ecosystem with an initial role as an accelerator providing a new layer of parallelism. Our approach will provide for co-design exploration towards the best QC accelerator choice for an application mix.\n\n\n\n\n\n\n\n\n\n\n\\clearpage\n\\appendices\n\n\\section*{Acknowledgment}\nThis research used the resources of the Argonne Leadership Computing Facility, which is a U.S. Department of Energy (DOE) Office of Science User Facility supported under Contract DE-AC02-06CH11357. We gratefully acknowledge the computing resources provided and operated by the Joint Laboratory for System Evaluation (JLSE) at Argonne National Laboratory. The authors would also like to acknowledge the NNSA's Advanced Simulation and Computing (ASC) program at Los Alamos National Laboratory (LANL) for use of their Ising D-Wave 2X quantum computing resource and D-Wave Systems Inc. for use of their 2000Q resource. The LANL research contribution has been funded by LANL Laboratory Directed Research and Development (LDRD). LANL is operated by Los Alamos National Security, LLC, for the National Nuclear Security Administration of the U.S. DOE under Contract DE-AC52-06NA25396. Clemson University is acknowledged for generous allotment of compute time on Palmetto cluster.\n\n\n\n\\bibliographystyle{.\/IEEEtranS}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} {"text":"\\section{Introduction and Motivation}\n\n\n This paper is motivated by work of Szlam, Maggioni, Coifman \\& Bremer \\cite{szlam2} and an observation made explicit by\nSzlam \\cite{szlam}: taking iterated spectral cuts induced by the nodal set of the first non-constant eigenfunction for the Neumann $p$-Laplacian seems to converge to rectangles in shape. \nIt is already observed in \\cite{szlam} that starting with an isosceles right triangle will lead to a spectral cut along the symmetry axis and produce two smaller isoceles right triangles: no convergence to rectangles takes place. However, this is unstable under small perturbations of the initial domain. \n\\begin{center}\n\\begin{figure}[ht!]\n\\includegraphics[width=0.85\\textwidth]{Fig1}\n\\caption{Iterated spectral cuts of the standard graph Laplacian seem to lead to rectangles in a generic, non-convex, non-smooth domain.}\n\\end{figure}\n\\end{center}\n\\vspace{-20pt}\n\nWe believe this to be a fascinating question in itself but an affirmative answer would also be useful in guaranteeing that iterative spectral partitioning is an effective method to partition domains and, ultimately, graphs and data (see Irion \\& Saito \\cite{saito} where this phenomenon is exploited). More importantly, a better understanding of this problem will shed light on the more general data analysis algorithms based on the $p$-Laplacian. More precisely, let $\\Omega \\subset \\mathbb{R}^2$ be an open, bounded, and connected domain. We now propose an iterative subdivision of $\\Omega$ as follows.\nFor any $1 \\leq p < \\infty$, the ground state of the $p-$Laplacian can be written as\n\\begin{equation} \\label{definition p Lap ground state}\n\\lambda_{1,p}(\\Omega) = \\inf_{\\int_{\\Omega}{f dx} = 0} \\frac{\\int_{\\Omega}{|\\nabla f|^p dx}}{ \\int_{\\Omega}{|f|^p dx}}.\\end{equation}\nIt is known that the function $f$ minimizing this functional exists and we can use it to iteratively define\n\\begin{equation}\n\\label{speccut}\n \\Omega_{n+1} = \\left\\{ x \\in \\Omega_n: f(x) \\geq 0\\right\\}\\,,\n \\end{equation}\nwhere $n=0,1,2,\\ldots$ and $\\Omega_0:=\\Omega$. \nThe function $f$ is only defined up to sign, so restricting to the part of the domain where it is positive is without loss of generality. We raise the following conjecture (and refer to the subsequent paragraphs for clarification and obvious obstacles).\n\\begin{quote} \\textbf{Main Conjecture.} If $\\Omega_0$ is not the isosceles right triangle (having angles 45-90-45), then the sequence of sets $(\\Omega_n)_{n=1}^{\\infty}$ converges to the set of rectangles with eccentricity\nbounded by 2 in the Gromov-Hausdorff distance.\n\\end{quote}\nIt is clear that one cannot expect convergence to a fixed rectangle: in general, the spectral cut of an $a \\times b$ rectangle with $a > b$ will be given by two $(a\/2) \\times b$ rectangles and, as long as $a \\leq 2b$,\nthe next cut would then yield two $(a\/2) \\times (b\/2)$ rectangles. This motivates a refined question.\n\n\\begin{quote} \\textbf{Question.} Do $(\\Omega_{2n})_{n=1}^{\\infty}$ and $(\\Omega_{2n+1})_{n=1}^{\\infty}$ converge in shape to a fixed rectangle?\n\\end{quote}\n\nThere is one obvious obstruction: if $\\Omega_{0}$ is not already a rectangle, then while performing iterated subdivisions, there is always a sequence of choices for the sign of the\neigenfunction that ensures that part of the boundary of $\\Omega_{n}$ coincides with part of the boundary of $\\Omega$ which could possibly be quite ill-behaved (say, fractal).\nHowever, for a sequence of choices of signs that leads to domains bounded away from $\\partial \\Omega_0$ ('deep' inside the domain) this should not be the case.\n\n\n\\begin{figure} \n\\centering\n\\includegraphics[width=.4\\textwidth]{Fig2-1}\\hspace{60pt}\n\\includegraphics[width=.4\\textwidth]{Fig2-2}\n\\caption{Left: the spectral cut of a quadrilateral determined by four corners, $(0, 0)$, $(\\pi\/25, 3\/5)$, $(1,0)$, and $(1, 3\/5-\\exp(1)\/100)$ provided by the graph 1-Laplacian. Right: the spectral cut of a shape close to a quadrilateral provided by the graph 1-Laplacian.}\\label{Fig:One}\n\\end{figure}\n\n\nWhile this particular question seems to be novel, the problem of trying to understand the geometry of nodal cuts induced by the $p-$Laplacian or general nonlinear operators has\nbeen studied for a long time. We refer to \\cite{plap,plap2,plap3,plap4,plap5,plap6,plap7,plap8,plap9,plap0,plap10,plap11} and references therein. We especially\nemphasize the works of Gajewski \\& G\\\"artner \\cite{plap7, plap8} who study the behavior of the cut as $p \\rightarrow 1$ as a means of finding effective ways of separating the\ndomain into two roughly equally sized domains, as well as the work of Parini \\cite{plap11} studying the limit of the cut as $p \\rightarrow 1$ under Dirichlet boundary conditions.\nMany of these results are posed for Dirichlet conditions where the effective functional as $p \\rightarrow 1$ is given by\n$$ \\inf_{E \\subset \\Omega}{ \\frac{ \\mathcal{H}^{n-1}(\\partial E)}{\\mathcal{H}^{n}(E)}} \\quad \\mbox{while the Neumann case induces} \\quad \\inf_{E \\subset \\Omega}{ \\frac{ \\mathcal{H}^{n-1}(\\partial E)}{\\mathcal{H}^{n}(E)\\mathcal{H}^{n}(\\Omega \\setminus E)}}.$$\nWhen the domain has Neumann boundary conditions, the quantity above is referred to as the {\\it Ratio Cut}.\nNew phenomena arise as a consequence. One interesting problem, that may also be of interest in the Dirichlet case, is the stability of the nodal set of the $p-$Laplacian\nas a function of $p$. \n\nIn the result presented here, we will focus on the $1$-Laplacian, where we can establish some preliminary results that suggest stability of rectangular domains as attractors for the proposed spectral dynamical system on domains. In particular, from henceforward, we refer to the {\\it $1$-spectral cut iteration} of a domain as the operation defined in \\eqref{speccut} with $p=1$ and the {\\it $1$-spectral cut} as the set $C_{n} := \\left\\{ x \\in \\Omega_n: f(x) = 0\\right\\}$. To simplify the terminology, we use {\\it spectral cut} and {\\it $1$-spectral cut} interchangeably. \n\n\\textbf{Numerics.} For the reader's convenience, we briefly recall here the numerical implementation of the p-Laplacian and its related spectral cut \\cite{Buhler_Hein:2009,Hein_Buhler:2010}.\nTake a point cloud $\\mathcal{X}:=\\{x_i\\}_{i=1}^N\\subset (\\mathcal M,d)$, a metric space $\\mathcal M$ with the metric $d$. Construct an affinity matrix ${\\bf W}\\in \\mathbb{R}^{N\\times N}$, where $\\bf W_{ij}$ is the affinity between $x_i$ and $x_j$. It is associated with an undirected affinity graph with $N$ vertices, where the affinity between $x_i$ and $x_j$, $w_{ij}$, is ${\\bf W}_{ij}$ for $i,j=1,\\ldots,N$. The {\\em graph $p$-Laplacian} is defined by\n\\begin{equation}\n\\Delta_p f(i)=\\sum_{j=1}^N w_{ij}\\phi_p(f(i)-f(j)),\n\\end{equation}\nwhere $f\\in\\mathbb{R}^N$ is a function defined on the vertices and $\\phi_p(x)=\\texttt{sign}(x)|x|^{p-1}$ for $x\\in \\mathbb{R}$. When $p=2$, this gives the bilinear form defined by the standard graph Laplacian ${\\bf D} - {\\bf W}$, where ${\\bf D}=\\texttt{diag}({\\bf W}{\\bf 1})$ and ${\\bf 1}$ is a $N$-dim vector with all entries $1$. Clearly, in general the graph p-Laplacian is nonlinear. We have\n\\begin{equation}\n\\langle f, \\Delta_p f \\rangle = \\frac12 \\sum_{i,j=1}^N w_{ij} |f_i - f_j|^p\n\\end{equation}\nfor $1 \\leq p < \\infty$. See for instance \\cite{Buhler_Hein:2009} for a discussion of the relationship between the variational formulation and the discrete operator formulation.\n\nA real number $\\lambda$ is called an eigenvalue for the graph $p$-Laplacian if there exists a non-zero vector $f\\in\\mathbb{R}^N$ so that \\cite[Definition 3.1]{Buhler_Hein:2009}\n\\begin{equation}\n(\\Delta_p f)_i=\\lambda \\phi_p(f_i)\\,,\n\\end{equation}\nwhere $i=1,\\ldots,N$. We call $f$ the $p$-eigenfunction of of the graph $p$-Laplacian associated with the eigenvalue $\\lambda$. We know that ${\\bf 1}$ is the trivial eigenvector with eigenvalue $0$, and we have \\cite[Lemma 3.2]{Buhler_Hein:2009}\n\\begin{equation}\n{\\bf 1}^T\\phi_p(f)=0\n\\end{equation} \nif the eigenvector $f$ is non-trivial.\nIt is shown in \\cite{Buhler_Hein:2009} that a non-zero function $f$ is an eigenvector if and only if it is a critical point (local minima) of the functional \n\\begin{equation}\nF_p(f)=\\frac{\\langle f, \\Delta_p f \\rangle}{\\|f\\|_{\\ell^p}^p}\\,.\n\\end{equation}\nNote that $F_p(f)$ is the discretization of the functional shown in \\eqref{definition p Lap ground state}.\nIn this work, our main interest is the second $p$-eigenvector of the graph $p$-Laplacian for the spectral clustering purpose. In \\cite{Buhler_Hein:2009}, the second eigenvalue is shown to be the global minimum of the functional\n\\begin{equation}\nF^{(2)}_p(f):=\\frac{\\langle f, \\Delta_p f \\rangle}{\\texttt{var}_p(f)},\n\\end{equation}\nwhere \n\\begin{equation}\n\\texttt{var}_p(f)=\\min_{c\\in\\mathbb{R}}\\sum_{i=1}^N |f_i-c|^p\\,,\n\\end{equation} \nand the corresponding eigenvector is then given by \n\\begin{equation}\nf_p^{(2)}=f^*-c^*{\\bf 1}, \n\\end{equation}\nwhere $f^*$ is any global minimizer of $F^{(2)}_p$ and $c^*=\\arg\\min_{c\\in \\mathbb{R}} \\sum_{i=1}^N |f^*_i-c|^p$. \nLike the usual spectral clustering, once we have $f_p^{(2)}$, we cluster the point cloud by the signs of its entries. \nFor $p=1$, \nthe consistency of spectral cut with the graph $1$-Laplacian was studied in \\cite{trillos2016consistency}.\nNumerically, we apply the nonlinear inverse power method proposed in \\cite{Hein_Buhler:2010} to evaluate the iterative bi-partition of a given 2-dimensional domain.\nIn Figure \\ref{Fig:One}, we present some numerically computed Ratio Cuts for nearly rectangular domains. \n\n\n\\textbf{Outline.} The paper proceeds as follows. In Section \\ref{sec:results}, we highlight the two main theorems we can prove on stability of near rectangular domains. We also present some open problems to be considered naturally as generalizations of these theorems. In Section \\ref{sec:pf1}, we give the full proof that the spectral cut algorithm converges to a rectangle with bounded aspect ratio if the initial domain is near a rectangle in Gromov-Hausdorff sense as will be carefully laid out below. Section \\ref{sec:pf3} provides the details for the proof that quadrilaterals near the rectangle of aspect ratio $2$ in terms of small angle deviations from $90$ degrees will converge under the spectral cut algorithm to rectangles with bounded aspect ratio. In the appendix we gather some long calculations that are useful in analyzing the Ratio Cut in a neighborhood of a quadrilateral.\n\n\n\n\n\\section*{Acknowledgements} \nWe thankfully acknowledge the generous support of NSF CAREER Grant DMS-1352353 (J.L. Marzuola \\& W. Hamilton). We thank Stefan Steinerberger for starting this project with us and for many helpful conversations along the way. \n\n\n\\section{results} \n\\label{sec:results}\nWe prove two results that, while not establishing the main conjecture, do seem to suggest a mechanism by which this procedure happens.\n$$ \\mbox{certain shapes} \\underbrace{\\implies}_{\\mbox{Theorem 1}} \\mbox{nearly straight cuts} \\implies \\mbox{curved quadrilaterals} \\underbrace{\\implies}_{\\mbox{Theorem 2}} \\mbox{rectangles.}$$\nThe missing steps are as follows: (1) we do not know whether a generic $\\Omega_0 \\subset \\mathbb{R}^2$ will ever produce domains $\\Omega_n$ for which Theorem 1 becomes applicable, for example, when $\\Omega_0$ has a fractal boundary; \nand (2) we do not know whether the dynamical system on the space of rectangles ever produces a quadrilateral sufficiently close to the set of rectangles for Theorem 2 to become applicable.\n\n\n\\subsection{Rectangular stability.} The first of the two main results states that the spectral cut of domains that 'roughly' look like rectangles are being cut\nin the middle. More formally, let us carefully describe the domains we will consider.\n\\begin{assumption}\n\\label{rectangle_assumptions}\nWe will work with domains that are small perturbations of a rectangle in the following sense:\n\\begin{itemize} \n\\item The domain $Q$ is a perturbed\nrectangle that is close in the Gromov-Hausdorff distance to a reference rectangle $R$: \n$$d_{GH}(Q, R) \\leq \\varepsilon \\left( \\mbox{length of the shorter side of}~R\\right)\\,,$$\nwhere $\\varepsilon>0$ is sufficiently small.\n\\item\nIn a roughly $10\\sqrt{\\varepsilon}-$neighborhood of the two intersection points of the $1$-spectral cut of $R$ with the boundary, the boundary of $Q$ \ncan be written as graphs of functions of the associated boundary segments of $R$ (see Fig. \\ref{fig:cut}). Moreover, each of these functions can be well approximated by a parabola with bounded, small curvature and small Lipschitz constant.\n\\end{itemize}\n\\end{assumption}\n\n\\begin{center}\n\\begin{figure}[ht!]\n\\begin{tikzpicture}[scale = 2]\n\\draw [] (0,0) -- (1.61,0) -- (1.61, 1) -- (0, 1) -- (0,0);\n\\draw [dashed] (0.8,-0.2) -- (0.8, 1.2);\n\\draw [ultra thick] (0,0) to[out=10,in=170] (1, 0);\n\\draw [ultra thick] (1,0) to[out=340,in=200] (1.61, 0);\n\\draw [ultra thick] (1.61,0) to[out=80,in=280] (1.61, 1);\n\\draw [ultra thick] (1.61,1) to[out=170,in=10] (0, 1);\n\\draw [ultra thick] (0,1) to[out=260,in=100] (0, 0);\n\\draw [ultra thick, dashed] (0.8, -0.2) to[out=85,in=277] (0.81, 1.2);\n\\end{tikzpicture}\n\\caption{A perturbed rectangle $Q$ close to a reference rectangle $R$: the longer part of the boundary of $Q$ can be written as the graph of a Lipschitz function. The goal is to show that the spectral cut of $Q$ is close to that of $R$.}\n\\label{fig:cut}\n\\end{figure}\n\\end{center}\n\n\\begin{theorem}\n\\label{thm1} \nFor a domain satisfying Assumption \\ref{rectangle_assumptions}, the spectral cut occurs in a $\\sqrt{\\varepsilon}-$neighborhood of the midpoint of the long axis. Moreover, there is a function $c(\\varepsilon)$ tending to\n0 as $\\varepsilon \\rightarrow 0$ such that, if said part of the boundary can be written as a Lipschitz function with Lipschitz constant $L \\leq \\sqrt{2} - c(\\varepsilon)$ in the $\\sqrt{\\varepsilon}-$neighborhood of the spectral cut of $R$, then the spectral cut of $Q$\nis a circular arc near a straight line.\n\\end{theorem}\n\nNumerical examples suggest that this result starts becoming applicable rather quickly: already a small number of cuts seems to suffice to produce roughly\nrectangular shapes. As soon as Theorem 1 becomes applicable the spectral cuts will start being closer and closer to straight lines, which immediately implies that many shapes will end\nup approaching quadrilaterals after several additional steps. The next result shows that as soon as we are dealing with quadrilaterals close to a rectangle, the procedure\nis smoothing and produces quadrilaterals closer to the rectangle; this has to be understood in the usual sense of 'cuts' away from the boundary: a quadrilateral\nhaving a $91^{\\circ}$ degree angle will always pass this angle on to one of its descendants. \nWe note a certain similarity of Theorem \\ref{thm1} to the result of Grieser-Jerison \\cite{GrieserJerison} for the standard Laplacian on rectangular domains of high aspect ratio with one side described by a curve. See also the recent work \\cite{BCM} where general curvilinear quadrilaterals were considered. \n\n\\subsection{Near Curvilinear Quadrilaterals} We settle our conjecture in the case of $\\Omega$ being near a quadrilateral that is close to a rectangle: we prove convergence to a rectangular shape in the Gromov-Hausdorff distance 'away from the boundary' (in the sense\ndiscussed above: rectangles incorporating parts of the initial boundary or boundaries created by very early initial spectral cuts need never be regular).\nWe introduce a qualitative way of measuring distance to quadrilaterals as follows: given a curvilinear quadrilateral $Q$ with angles $\\alpha, \\beta, \\gamma, \\delta$ and sides described by the curves $y = \\gamma_1 (x)$, $y = \\gamma_2(x)$, $x = \\gamma_3(y)$ and $x = \\gamma_4 (y)$, we\ndefine the functional\n\\begin{equation}\\label{Definition I(Q)}\nI(Q) = \\left| \\alpha- \\frac{\\pi}{2} \\right| + \\left| \\beta- \\frac{\\pi}{2} \\right| + \\left| \\gamma- \\frac{\\pi}{2} \\right| + \\left| \\delta- \\frac{\\pi}{2} \\right| + \\max_{1 \\leq j \\leq 4} \\sup_s (|\\gamma_j' |(s) + |\\gamma_j'' |(s)+ |\\gamma_j''' |(s)) .\n\\end{equation}\nWe particularly want our domains to be well approximated by parabolic curves on the sides intersecting the ratio cut, and will refer to such domains as {\\it approximately parabolic curvilinear quadrilaterals}. \n\n\\begin{remark}\nThis is actually quite a bit more restrictive than one needs, though is certainly sufficient; otherwise, the theorem becomes harder to frame as one must introduce a large number of cases for behaviors of each boundary component of the domain. However, in the calculations below, we will work with domains that are perturbations of a base rectangle with a reasonable aspect ratio, and whose top and bottom boundary components can be well-approximated by parabolas in a neighborhood near the axis of symmetry of the base rectangle. As a result, we can dramatically change regularity requirements for the left and right curves in our model calculations as long as the perturbations are nearly symmetric in area and bounded by a sufficiently small constant. \n\\end{remark}\n\n\\begin{theorem} \n\\label{thm2}\nFor an approximately parabolic curvilinear quadrilateral of aspect ratio near $1:2$, there exists $\\varepsilon_1 > 0$ such that if $I(Q) \\leq \\varepsilon_1$, then the $1$-spectral cut induced\nby the $L^1-$Laplacian is a circular arc with opening angle bounded above by $\\varepsilon_1\/8$. \n\\end{theorem}\nThe proof also shows that the constant $1\/8$ cannot be further improved.\nThis result implies that near-quadrilateral regions have $1$-spectral cuts that are closer to a quadrilateral in Gromov-Hausdorff distance. This implies exponentially fast convergence to rectangles in shape. The way the result is obtained actually allows for a fairly precise understanding of what happens (in particular, it can be used to show that there is a choice of signs such that $I(Q_n)$ grows substantially along the subsequence). We refer to the proof for details. \n\n\n\\subsection{Open problems} These results naturally suggest several open problems; we only name a few that seem \nparticularly natural.\n\n\\begin{enumerate}\n\n\\item Prove the main conjecture for $p=1$; that is, show that a generic $\\Omega_0 \\subset \\mathbb{R}^2$ will produce domains $\\Omega_n$ for which Theorem 1 can be applied, and show that the dynamical system on the space of rectangles produces a quadrilateral sufficiently close to the set of rectangles for Theorem 2 to become applicable. Is it possible to transfer some of the arguments to the range $p \\in (1,1+\\varepsilon_0)$? What happens\nas $p \\rightarrow \\infty$? \n\\item What happens in higher dimensions, or even to domains with curvature, like manifolds? One would still expect the cut to have a smoothing effect but the types of geometric obstruction that\none could encounter in the process may be more complicated. Is the generic limit given by a rectangular box? If not, is there a finite set of shapes\nthat can arise in the limit?\n\\item It seems natural to expect that similar results should hold true under Dirichlet conditions. This would require the study of the variational problem \n$$ \\inf_{E \\subset \\Omega}{ \\frac{ \\mathcal{H}^{n-1}(\\partial E)}{\\mathcal{H}^{n}(E)}}.$$\nCan the results be extended to this case?\n\\item Experimentally, we observe that the nodal line is (generically) fairly stable under small perturbations of the value of $p$. Can any result in this\ndirection be made precise? Does it help to assume the domain $\\Omega$ to be convex?\n\\end{enumerate}\n\n\n\n\n\\section{proof of theorem 1}\n\\label{sec:pf1}\n\n\\subsection{Preliminaries.} The crucial ingredient in our approach is that the $p$-Laplacian degenerates as $p \\rightarrow 1^+$. The minimum of the associated energy functional characterizing the ground state is not assumed by any continuous\nfunction: reinterpreting the functional in terms of total variation, the extremal function is constant on two sets in the domain that are separated by\nan $(n-1)-$dimensional hypersurface. More precisely, we have the following consequence of the coarea formula (see \\cite{stein} for further implications\nof this result).\n\n\n\\begin{thm}[Cianchi, \\cite{Cianchi}] Let $\\Omega \\subset \\mathbb{R}^{n}$ be open, bounded and connected. Then we have the sharp Poincar\\'e-type inequality for any sufficiently smooth functions $u$:\n$$\\left\\|u-\\frac{1}{|\\Omega|}\\int_{\\Omega}{u(z)dz}\\right\\|_{L^{1}(\\Omega)} \\leq\n \\left(\\sup_{\\Gamma \\subset \\Omega}{\\frac{2}{\\mathcal{H}^{n-1}(\\Gamma)}\\frac{|S||\\Omega \\setminus S|}{|\\Omega|}}\\right)\\left\\|\\nabla u\\right\\|_{L^{1}(\\Omega)},$$\nwhere $\\Gamma \\subset \\Omega$ ranges over all surfaces which divide $\\Omega$ into two connected open subsets $S$ and $\\Omega \\setminus \\overline S$. \n\\end{thm}\n\nThis means that the nodal line $\\Gamma$ defining the $1$-spectral cut is a hypersurface partitioning $\\Omega$ into two sets $S, \\Omega \\setminus \\overline S$\n$$ \\mbox{so as to minimize the quantity} \\qquad \\frac{\\mathcal{H}^1(\\Gamma) |\\Omega|}{|S| |\\Omega \\setminus S|}.$$\nIt is relatively easy to check that this value is assumed if we relax the $L^1$-norm of the gradient and re-interpret it as total variation. Let us assume that $f(x) = a \\chi_{S} + b \\chi_{\\Omega \\setminus S}$.\nWe want $f$ to have mean value 0, which leads to\n$$ a |S| + b |\\Omega \\setminus S| = 0 \\quad \\mbox{and thus} \\quad b = -a\\frac{|S|}{|\\Omega \\setminus S|} \\quad \\mbox{implying} \\quad \\|f\\|_{L^1} = 2a |S|,$$\nwhere we take $a>0$ without loss of generality.\nMoreover, the ``formal'' contribution to the total variation interpretation of the gradient is given by\n$$ \\|\\nabla f \\|_{L^1} = |\\Gamma| (a - b) =|\\Gamma| \\left(a + a\\frac{|S|}{|\\Omega \\setminus S|} \\right).$$\nThis implies\n$$ \\frac{ \\|\\nabla f \\|_{L^1} }{ \\|f\\|_{L^1}} = \\frac12 |\\Gamma| \\left( \\frac{1}{|S|} + \\frac{1}{|\\Omega\\setminus S|}\\right) = \\frac12 |\\Gamma| \\frac{|\\Omega|}{|S| |\\Omega \\setminus S|}.$$\nTo be rigorous, convolution with a smooth mollifier shows that one can get arbitrarily close to the optimal constant with smooth functions. We will now show that, for a rectangle, the optimal spectral cut splits the domain via a straight line cut intersecting the longest sides at right angles. The crucial ingredient here is that the argument does not appeal to symmetry,\nis stable under perturbations and easily implies the same results for domains that are merely close to rectangles in the Gromov-Hausdorff distance.\n\n\\begin{lemma} \\label{Lemma: rectangle result}\nLet $R = [0,a] \\times [0,b] \\subset \\mathbb{R}^2$ with $a > b$. The $1$-spectral cut is exactly at $\\left\\{a\/2\\right\\} \\times [0,b]$.\n\\end{lemma}\n\\begin{proof}\nThe cut in the middle of the longer side yields\n$$ \\frac{\\mathcal{H}^1(\\Gamma) |\\Omega|}{|S| |\\Omega \\setminus S|} = \\frac{b (ab)}{(ab)^2\/4} = \\frac{4}{a}.$$\nIt is clear that other spectral cuts touching the longer sides have $ \\frac{\\mathcal{H}^1(\\Gamma) |\\Omega|}{|S| |\\Omega \\setminus S|} \\geq \\frac{4}{a},$ since $\\mathcal{H}^1(\\Gamma)\\geq b$ and $|S| |\\Omega \\setminus S|\\leq (ab)^2\/4$. \n\\begin{center}\n\\begin{figure}[ht!] \n\\begin{tikzpicture}[scale = 2]\n\\draw [ultra thick] (0,0) -- (1.61,0) -- (1.61, 1) -- (0, 1) -- (0,0);\n\\draw [dashed] (0.8,0) -- (0.8, 1);\n\\draw [thick] (0.7,0) to[out=19,in=140] (1.61, 0.6);\n\\node at (1.5,0.2) {$S$};\n\\node at (0.3,0.6) {$\\Omega \\setminus S$};\n\\end{tikzpicture}\n\\caption{The geometric construction in the proof of Lemma 1. }\n\\label{fig4}\n\\end{figure}\n\\end{center}\n\nIf $\\Gamma$ touches two opposite sides, then\n$$ \\frac{\\mathcal{H}^1(\\Gamma) |\\Omega|}{|S| |\\Omega \\setminus S|} \\geq b \\frac{ |\\Omega|}{|S| |\\Omega \\setminus S|} \\geq \\frac{b (ab)}{(ab)^2} \\min_{0 \\leq x \\leq 1}{\\frac{1}{x(1-x)}} \\geq \\frac{4}{a}.$$\nEquality can only arise if the cut has length $b$ (forcing it to be a line) and $x=1\/2$ (forcing a split into two domains of the same area). This characterizes\nexactly the cut in the middle.\n\n\nIt remains to deal with the case where the spectral cut touches two adjacent sides; this case is illustrated in Figure \\ref{fig4}. Let us assume that the enclosed domain is denoted by $S$, $|S| = x |\\Omega| = x ab$ for some $0 < x < 1$, and $\\Gamma = \\partial S \\cap \\mbox{int}~R$ is the part of the boundary curve strictly inside the rectangle.\nThe isoperimetric inequality implies that\n$$ \\mathcal{H}^1(\\Gamma) \\geq \\sqrt{\\pi}\\sqrt{x a b},$$\nand thus\n$$ \\frac{\\mathcal{H}^1(\\Gamma) |\\Omega|}{|S| |\\Omega \\setminus S|} \\geq \\frac{ \\sqrt{\\pi x a b} ab }{a^2 b^2 x (1-x)} = \\frac{\\sqrt{\\pi}}{\\sqrt{x}(1-x)} \\frac{1}{\\sqrt{a}\\sqrt{b}}.$$\nAn explicit computation shows that for all $0 < x < 1$\n$$ \\frac{1}{\\sqrt{x}(1-x)} \\geq \\frac{3\\sqrt{3}}{2}$$\nand therefore, using $b \\geq a,$\n$$ \\frac{\\mathcal{H}^1(\\Gamma) |\\Omega|}{|S| |\\Omega \\setminus S|} \\geq \\frac{3\\sqrt{3 \\pi}}{2} \\frac{1}{a} \\geq \\frac{4.6}{a}.$$\nThis is always a constant fraction worse than the cut along the longest axes, concluding the argument.\n\\end{proof}\n\n\\begin{remark}\n\\label{rmk1}\nIt is easily seen that all aspects of the argument are stable under (even moderately large) perturbations of the domain in the sense of Gromov-Hausdorff distance. To be specific, take a domain $Q$ satisfying Assumption \\ref{rectangle_assumptions}, and denote the length of the shorter side of the associated rectangle $R$ by $h$. By Assumption \\ref{rectangle_assumptions}, the Gromov-Hausdorff distance between $Q$\nand $R$ is $d_{GH}(Q, R) \\leq \\varepsilon h$. There is clearly a straight line that is parallel to the shorter side of $R$ and bisects the area: the\nbound on the Gromov-Hausdorff distance implies that this straight line has length at most $h + 2h\\varepsilon$. Therefore,\n$$ \\inf_{\\Gamma}{\\frac{\\mathcal{H}^1(\\Gamma)}{|S| |\\Omega \\setminus S|}} \\leq 4h + 8h \\varepsilon.$$\nArguments as in Lemma 1 show that it has to cut the longer side roughly in the middle. \n\\end{remark}\n\n\n\nThe second ingredient is a straightforward consequence of the isoperimetric inequality that we need to make quantitative.\n\\begin{center}\n\\begin{figure}[ht!]\n\\begin{tikzpicture}[scale = 6]\n\\draw [ultra thick] (0,0) -- (1,0);\n\\node at (-0.02, -0.04) {$a$};\n\\node at (1.02, -0.04) {$b$};\n\\node at (0.5, 0.05) {$\\Omega$};\n\\draw [ultra thick] (0,0) to[out=20,in=180] (0.4, 0.08);\n\\draw [ultra thick] (0.4,0.08) to[out=20,in=160] (1, 0);\n\\end{tikzpicture}\n\\caption{A curve slightly longer than the straight line.}\n\\label{fig:areagain}\n\\end{figure}\n\\end{center}\n\n\n\n\n\\begin{lemma} Consider a simply connected domain $\\Omega$ as shown in Figure \\ref{fig:areagain} that is comprised of a straight line segment between $a$ and $b$ and some arbitrary curved line segment enclosing a domain $\\Omega$. For every $\\varepsilon_0 >0$\nthere exists an $\\varepsilon_1 > 0$ such that if\n\\begin{equation}\n\\label{assump1}\n|\\partial \\Omega| \\leq \\left(1+\\varepsilon_1\\right) 2\\|a - b\\|,\n\\end{equation}\nthen\n$$ |\\Omega| \\leq \\frac{1 + \\varepsilon_0}{\\sqrt{6}} \\| a - b\\|^{3\/2} \\sqrt{|\\partial \\Omega| - 2 \\|a-b\\|}.$$\n\\end{lemma}\n\\begin{proof} The isoperimetric inequality immediately implies that the optimal curve defining the Ratio Cut for $Q$ satisfying Assumption \\ref{rectangle_assumptions} must take the form of a circular arc; otherwise, we could fix $a,b$ at the boundary of a circle\nand use the novel shape to create a set in the plane that contains more area than the disk whose boundary has the same size. \n\n It remains to compute the constants for the circular arc.\nChoose a coordinate system with $a,b$ at $(0,0),(\\|a-b\\|,0)$ respectively. The opening angle $\\alpha$ of the sector of the disk with radius $r>0$ that produces such an arc is given by\n$$ \\alpha = 2 \\arcsin{\\left( \\frac{\\|a-b\\|}{2r} \\right)},$$\nwhere $0 < \\alpha < \\frac{\\pi}{2}$ is sufficiently small by assumption \\eqref{assump1}.\nAs a consequence, the length of the circular arc is given by\n$$ r \\alpha = 2r \\arcsin{\\left( \\frac{\\|a-b\\|}{2r}\\right)} \\geq \\|a-b\\| + \\frac{\\|a-b\\|^3}{24} \\frac{1}{r^2},$$\nand so\n\\begin{equation}\n\\label{pomegabd}\n |\\partial \\Omega| \\geq 2\\|a-b\\| + \\frac{\\|a-b\\|^3}{24} \\frac{1}{r^2}.\n \\end{equation}\nIn particular, by assumption \\eqref{assump1}, we have that \n$$ \\frac{\\|a-b\\|^3}{24} \\frac{1}{r^2} \\leq 2\\varepsilon_1\\|a-b\\| \\qquad \\mbox{and thus} \\qquad \\frac{\\|a-b\\|^2}{r^2} \\leq 48 \\varepsilon_1.$$\n\nThe enclosed area captured by the line segment and the circular arc is \n$$\\frac{ r^2 \\pi \\alpha}{2\\pi} - r^2 \\cos{\\left(\\frac{\\alpha}{2}\\right)} \\sin{\\left(\\frac{\\alpha}{2}\\right)}.$$\nWe have\n$$ \\frac{ r^2 \\pi \\alpha}{2\\pi} = r^2 \\arcsin{\\left( \\frac{\\|a-b\\|}{2r} \\right)} = r \\frac{\\|a-b\\|}{2} + \\frac{\\|a-b\\|^3}{48 r} + \\mbox{higher order terms}$$\nas well as\n$$ \\cos{\\left(\\frac{\\alpha}{2}\\right)} = \\sqrt{1 - \\frac{\\|a-b\\|^2}{4r^2}} \\quad \\mbox{and} \\quad \\sin{\\left(\\frac{\\alpha}{2}\\right)} = \\frac{\\|a-b\\|}{2r}.$$\nThe higher order terms in the expansion are an infinite series in $\\|a-b\\|\/r$ with exponents decaying fast enough. For $\\varepsilon_1$ sufficiently\nsmall depending on $\\varepsilon_0$, we can bound\n$$ \\mbox{higher order terms} \\leq \\varepsilon_0 \\frac{\\|a-b\\|^3}{48 r} .$$\nAltogether this implies that\n$$\\mbox{area} \\leq (1 + \\varepsilon_0) \\frac{\\|a-b\\|^3}{12r}$$\nand plugging in the bound on $1\/r$ from \\eqref{pomegabd} gives\n$$\\mbox{area} \\leq \\frac{1 + \\varepsilon_0}{\\sqrt{6}} \\| a - b\\|^{3\/2} \\sqrt{|\\partial \\Omega| - 2 \\|a-b\\|}.$$\n\\end{proof}\n\n\n\n\n\n\\subsection{Proof of Theorem 1}\n\n\\begin{proof} \nWe assume without loss of generality $|Q| = 1$. For this proof we minimize, over all possible set partitions $Q = S \\cup (Q \\setminus S)$, the quotient\n$$ \\inf_{\\Gamma} \\frac{\\mathcal{H}^1(\\Gamma) }{|S| | Q \\setminus S|},$$\nwhere $\\Gamma=\\overline S\\cap \\overline{Q \\setminus S}$.\n\nLemma \\ref{Lemma: rectangle result} and Remark \\ref{rmk1} show that the cut must intersect opposite sides; if the aspect ratio is large enough, then these opposite sides will be the longest sides. \n\n\n\n\n\nWe now show that the distance to the axis of symmetry of $R$ is of order $\\lesssim \\sqrt{\\varepsilon}$.\nSince the Gromov-Hausdorff distance between the domain and the rectangle is $\\varepsilon$, by the same calculation in Remark \\ref{rmk1} we see that any cut has to have length at least\n$ h (1- 2\\varepsilon)$, where $h$ is the length of the shorter side of the associated rectangle $R$. Note that the shortest line between two points may not make the optimal Ratio Cut, though, as that may result in a less favorable area splitting. Hence, one possibility is that we get a gain from having a slightly longer curve to encompass a more favorable area, but then by the isoperimetric inequality, it can be seen that circular arcs are favorable to any other configuration in terms of length to area trade-offs. A key example of this is the trapezoidal domain with angled top and bottom boundary curves. Therefore the optimal spectral cut $Q = A \\cup B$ satisfies\n$$ \\frac{h(1-2\\varepsilon)}{|A| |B|} = \\frac{h(1-2\\varepsilon)}{|A| (1-|A|)} \\leq \\frac{\\mathcal{H}^1(\\Gamma)}{|S| |Q \\setminus S|} \\leq 4h(1 + 2 \\varepsilon)$$\nfor any $\\Gamma$.\nThus, for $\\varepsilon \\leq 0.1$\n\\begin{equation}\n\\frac{1}{|A| (1-|A|)} \\leq 4\\frac{1 + 2\\varepsilon}{1 - 2 \\varepsilon} \\leq 4 + 20\\varepsilon.\\label{bound of A(1-A)}\n\\end{equation}\nWhen combined with the elementary inequality\n$$ \\frac{1}{x(1-x)} \\geq 4 + 16\\left(x-\\frac{1}{2}\\right)^2,$$\nwe have that the optimal spectral cut yields two sets satisfying\n$$ \\frac{1}{2} - \\sqrt{\\frac{5\\varepsilon}{4}} \\leq |A| \\quad \\mbox{and}\\quad |B| \\leq \\frac{1}{2} + \\sqrt{\\frac{5\\varepsilon}{4}}.$$\nThe Lipschitz bound then ensures that the spectral cut's intersection with $\\partial Q$ occurs in a $\\sqrt{\\epsilon}$ small neighborhood of $Q$'s axis of symmetry. \n\nSince the functional itself only contains the length $\\mathcal{H}^1(\\Gamma),$ as well as a term depending on the partition of the areas $|S||\\Omega\\setminus S|$, we can conclude that the optimal $\\Gamma$ is a circular arc; otherwise, we can minimize the length of the curve while keeping the bounded area fixed. Indeed, this boils down to the question of minimizing the arc-length of a curve of fixed area,\n\\[\n\\text{arg min}_{y = \\gamma (x)} \\left\\{ \\int_a^b \\sqrt{1 + (\\gamma'(x) )^2} dx \\ \\bigg| \\ \\int_a^b |\\gamma| dx \\,\\,\\mbox{ fixed.}\\right\\},\n\\] \nwhich is minimized by a circular arc via the isoperimetric inequality.\nWe already know from above that $\\mathcal{H}^1(\\Gamma) \\leq (1+2\\varepsilon) h$, so Lemma 2 implies that the area captured by the curved\narc satisfies\n$$ \\mbox{captured area} \\leq \\frac{1 + c_{\\varepsilon}}{\\sqrt{6}} h^{3\/2} \\sqrt{\\mathcal{H}^1(\\Gamma) - h} \\,\\leq \\frac{1 + c_{\\varepsilon}}{\\sqrt{3}} h^2 \\sqrt{\\varepsilon} ,$$\nwhere $c_{\\varepsilon} > 0$ depends on $\\varepsilon$, but does tend to 0 as $\\varepsilon$ tends to $0$. \n\nIn summary, the above conclusions show that the Ratio Cut $\\Gamma$ must be a circular arc in a $\\sqrt{\\epsilon}$ neighborhood of the axis of symmetry of the reference rectangle from Assumption $1$. \n\\end{proof}\n\n\n\\section{Proof of Theorem \\ref{thm2} for Curvilinear Quadrilaterals with Parabolic Top and Bottom Curves}\n\\label{sec:pf3}\n\nTo prove Theorem $2$, we will first analyze how the Ratio Cut depends on parameters determining the domain $Q$ and circular arc cut $\\Gamma$ for a true quadrilateral. Then, we will demonstrate for a simple example of a curvilinear trapezoid with one parabolic edge and three flat edges that the curvature has a smaller order impact on the Ratio Cut than the trapezoidal feature. Lastly, we will prove the theorem for even more general domains with parabolic top and bottom bounding curves. These domains will be shown to be generic in section 5, at least to leading order in how the Ratio Cut depends upon the structure of the curves that make up the longest sides of our approximately parabolic curvilinear rectangles. \n\nConsider the quadrilateral $Q$ determined by the vertices\n$$ (0,0), (x_1, a), (1,0) \\quad \\mbox{and} \\quad (1+x_2, a+x_3),$$\nwhere we assume that $ 0 < a < 1$ (implying that we have normalized the quadrilateral and put the longer side on the $x-$axis). The quantities $x_1, x_2, x_3$\nare perturbation parameters and assumed to be small, in the sense that $|x_i| \\ll \\min\\left\\{a,1- a\\right\\}$. The condition $|x_i| \\ll a$ is clear; we want the perturbation to be small with respect to length and height of the rectangle. The other condition is slightly more subtle: if it is not satisfied, then the quadrilateral might be close to a square and the location of the spectral cut will depend nonlinearly on the perturbation parameters. The cut is still going to be a circular arc nearly bisecting the domain, but a slight perturbation of $x_1, x_2, x_3$ may send $\\Gamma$ from being nearly vertical to nearly horizontal (see Fig. \\ref{fig:quad}).\nThe same arguments used to prove Theorem 1 show that the spectral cut will be a circular arc connecting two points ${\\bf q} = (q,0)$ and ${\\bf p} = (p,y(p))$.\n\n\n\\begin{center}\n\\begin{figure}[ht!]\n\\begin{tikzpicture}[scale=4]\n\\draw [ultra thick] (0,0) -- (1,0) -- (1.1, 0.8) -- (-0.1, 0.7) -- (0,0);\n\\filldraw (0,0) circle (0.015cm);\n\\node at (-0.08, -0.08) {(0,0)};\n\\node at (0.08, 0.08) {$\\alpha$};\n\\filldraw (1,0) circle (0.015cm);\n\\node at (0.95, 0.05) {$\\beta$};\n\\node at (1.08, -0.08) {(1,0)};\n\\filldraw (1.1,0.8) circle (0.015cm);\n\\node at (1.2, 0.8+0.08) {($1+x_2$,$a+x_3$)};\n\\filldraw (-0.1,0.7) circle (0.015cm);\n\\node at (-0.24, 0.8-0.08) {($x_1$,$a$)};\n\\filldraw (0.5,0) circle (0.015cm);\n\\node at (0.5, -0.08) {${\\bf q}:=(q,0)$}; \n\\filldraw (0.45,0.75) circle (0.015cm);\n\\node at (0.48,0.82) {${\\bf p}:=(p,y (p)) $}; \n\\draw [thick,dashed] (0.45, 0.75) -- (0.5,0);\n\\draw [thin,dashed] (0.5,0) to[out=45,in=330] (0.45, 0.75);\n\\node at (0.42, 0.35) {$\\Gamma$};\n\\node at (0.69, 0.35) {$\\tilde{\\Gamma}$};\n\\node at (1, 0.7) {$\\gamma$};\n\\node at (-0.03, 0.63) {$\\delta$};\n\\node at (0.43, 0.05) {$\\eta$};\n\\node at (0.55, 0.1) {$\\nu$};\n\\node at (0.4, 0.67) {$\\phi$};\n\\node at (0.51, 0.65) {$\\mu$};\n\\end{tikzpicture}\n\\caption{A general quadrilateral $Q$ after rescaling.}\n\\label{fig:quad}\n\\end{figure}\n\\end{center}\n\n\n\\subsection{Circular Arc to Triangle in terms of the sector angle}\n\n\nLet us assume the cut, $\\tilde E$, is a circular arc from $\\bf p$ to $\\bf q$, where $\\bf p$ and $\\bf q$ appear nearly opposite each other on the longest sides. We will take $|\\tilde \\Gamma| = r \\theta$ for some radius $r>0$ and sector angle $\\theta\\geq0$, to be determined, and observe then the area contained between the line $\\Gamma$ connecting $\\bf p$ and $\\bf q$ and the circular arc $\\tilde \\Gamma$, which we will call $\\tilde \\Omega$, satisfies\n\\[\n| \\tilde \\Omega | = \\frac{r^2 \\theta}{2} - \\frac{\\| {\\bf p}- {\\bf q}\\|^2}{4 \\tan (\\theta\/2)}.\n\\]\nWe also have that\n\\[\nr = \\frac{\\| {\\bf p}- {\\bf q} \\|}{2 \\sin (\\theta\/2)},\n\\]\nfrom which we conclude, after Taylor expanding in $\\theta$, that\n\\[\n| \\tilde \\Omega | = \\|{\\bf p}- {\\bf q}\\|^2 \\left( \\frac{\\theta^2}{48} + O( \\theta^4) \\right).\n\\]\nAlso, Taylor expanding once more, we have\n\\[\n|\\tilde \\Gamma| = r \\theta = \\| {\\bf p}- {\\bf q}\\| \\left( 1 + \\frac{\\theta^2}{24} + O( \\theta^4) \\right).\n\\]\n\nWith these identities in hand, we can proceed to explore how the ratio cut depends upon the curves defining the longest aspect ratio sides of our curvilinear quadrilateral. To illustrate how to compute the dependence of the cut locally on the parametrization of a curve, we will first work with a toy model that is flat on $3$ sides and has a smooth quadratic curve on one of the long sides. \n\n\\subsection{The parabolic trapezoid}\n\\label{partrap}\n\nWe take a domain that is a parabolic trapezoid bounded by a straight line from $(0,0)$ to $(1,0)$, a straight line from $(0,0)$ to $(0,1\/2)$, the straight line from $(1,0)$ to $(1, 1\/2+a)$ and the curve $y(x) = \\epsilon x^2 + (a - \\epsilon) x + 1\/2$; note that, for convenience, we have chosen our aspect ratio to be approximately $2$. Our goal is to track how the cut changes as a function of $\\epsilon,a$ close to $0$. The Ratio Cut will intersect the bottom and top boundary curves at the points $(q,0)$ and $(p,y(p))$ respectively, with the cut itself a circular arc. Splitting the domain by a circular arc, we can decompose one of the cut domains into two curvilinear quadrilaterals, see Figure \\ref{fig:quad2} for an illustration.\n\n\\begin{center}\n\\begin{figure}[ht!]\n\\begin{tikzpicture}[scale=4]\n\\draw [ultra thick] (0,1\/2)--(0,0) -- (1,0) -- (1, 1\/2+0.1) \n\\draw [ultra thick] (0,0.5) to[out=25,in=160] (1, 0.6);\n\\filldraw (0,0) circle (0.015cm);\n\\node at (-0.08, -0.08) {(0,0)};\n\\filldraw (1,0) circle (0.015cm);\n\\node at (1.08, -0.08) {(1,0)};\n\\filldraw (1,1\/2+0.1) circle (0.015cm);\n\\node at (1.08, 1\/2+0.1+0.08) {($1$,$0.6$)};\n\\filldraw (0,0.5) circle (0.015cm);\n\\node at (-0.1, 0.5+0.08) {($0$,$0.5$)};\n\\filldraw (0.5,0) circle (0.015cm);\n\\node at (0.5, -0.08) {${\\bf q}:=(q,0)$}; \n\\filldraw (0.52,0.67) circle (0.015cm);\n\\node at (0.48,0.82) {${\\bf p}:=(p,y (p)) $}; \n\\draw [thin,dashed] (0.5,0) to[out=75,in=285] (0.52, 0.67);\n\\node at (0.45, 0.35) {$\\Gamma$};\n\\end{tikzpicture}\n\\caption{A parabolic trapezoid $Q$ as a simplified model.}\n\\label{fig:quad2}\n\\end{figure}\n\\end{center}\n\nIt can be easily seen that to compute the Ratio Cut, we can split the domains into components given by the region $[0,p]$, a triangle with base $[p,q]$, and the cap of a circular arc that is either added or subtracted depending upon orientation ($p 0$ or $\\theta < 0$ respectively). Decomposing the domain into these components, we compute\n\\begin{align*}\n&A_{total} = \\int_0^1 y(x) dx = \\frac{\\epsilon}{3} + \\frac{a-\\epsilon}{2} + \\frac12, \\ \\ \\text{(Total Area)} \\\\\n&A_1 (p,q,\\theta) = \\int_0^p y(x) dx + \\frac12 y(p) (q-p)+\\frac12 R^2 (p,q,\\theta) \\theta \\\\\n& \\hspace{4cm} - \\frac14 \\left[ (q-p)^2 + y(p)^2 \\right] \\cot \\left( \\frac{\\theta}{2} \\right), \\ \\ \\text{(Left Area)} \\\\\n& A_2 (p,q,\\theta) = A_{total} - A_1 (p,q,\\theta), \\ \\ \\text{(Right Area)} \n\\end{align*}\nwhere $R(p,q,\\theta)$ is the radius of the circle, given by\n\\begin{align}\nR(p,q,\\theta) = \\frac{\\sqrt{ (q-p)^2 + y(p)^2 }}{ 2 \\sin \\left( \\frac{\\theta}{2} \\right)}\\,.\n\\end{align}\nThe above quantities give the ratio cut, denoted by $\\text{RC} (p,q,\\theta)$:\n\\begin{align*}\n \\text{RC} (p,q,\\theta) = \\frac{ R(p,q,\\theta) \\theta}{ A_1 (p,q,\\theta) A_2 (p,q,\\theta)}.\n\\end{align*}\nIt follows via direct calculation that $ \\text{RC} (p,q,\\theta)$ is a smooth function of $(p,q,\\theta)$ in a neighborhood of $(1\/2,1\/2,0)$. Indeed, since $\\frac12 R^2 (p,q,\\theta) \\theta - \\frac14 \\left[ (q-p)^2 + y(p)^2 \\right] \\cot \\left( \\frac{\\theta}{2} \\right)\\to 0$ and $R(p,q,\\theta) \\theta\\to \\sqrt{ (q-p)^2 + y(p)^2 }$ when $\\theta\\to 0$, $\\text{RC} (p,q,\\theta)$ is smooth in a neighborhood of $(1\/2,1\/2,0)$. Hence we can explore behaviors nearby using the Implicit Function Theorem. \n\nAs an illustrative calculation, let us simply take the full quadratic approximation in $\\theta$, $p-\\frac12$, $q-\\frac12$, $a$ and $\\epsilon$ to the Ratio Cut. A direct calculation gives:\n\\begin{small}\n\\begin{align*}\n& \\text{RC} (p,q,\\theta) {\\color{red}=} \\left( 8 + 24 \\left(q - \\frac12 \\right)^2 - 16 (q-\\frac12) (p-\\frac12) + 24 (p - \\frac12)^2 \\right) \\\\\n& - a \\left(8 + 8\\left(q - \\frac12 \\right) - 8 \\left(p - \\frac12 \\right) - 56\\left(q - \\frac12 \\right) ^2 - 56 \\left(p - \\frac12 \\right)^2 + 80 \\left(q - \\frac12 \\right) \\left(p - \\frac12 \\right) \\right) \\\\\n& + \\epsilon \\left( \\frac43 + \\frac{52}{3} \\left(q - \\frac12 \\right)^2 + \\frac{100}{3} \\left(p - \\frac12 \\right)^2 - 40 \\left(q - \\frac12 \\right) \\left(p - \\frac12 \\right) \\right) + \\frac43 \\theta \\left( \\left(q - \\frac12 \\right) + \\left(p - \\frac12 \\right) \\right) \\\\\n& + a^2 \\left( 10 + 16 \\left(q - \\frac12 \\right) - 16 \\left(p - \\frac12 \\right) + 120 \\left(q - \\frac12 \\right)^2 + 104 \\left(p - \\frac12 \\right)^2 -192 \\left(q - \\frac12 \\right) \\left(p - \\frac12 \\right) \\right) \\\\\n& + \\epsilon^2 \\left( \\frac{122}{9} \\left(q - \\frac12 \\right)^2 + \\frac{218}{9} \\left(p - \\frac12 \\right)^2 - \\frac{284}{9} \\left(q - \\frac12 \\right) \\left(p - \\frac12 \\right) \\right) + \\frac{7}{18} \\theta^2 \\\\\n& -a \\epsilon \\left( \\frac83 + \\frac{8}{3} \\left(q - \\frac12 \\right) - 8 \\left(p - \\frac12 \\right) + 72 \\left(q - \\frac12 \\right)^2 + 104 \\left(p - \\frac12 \\right)^2 - \\frac{464}{3} \\left(q - \\frac12 \\right) \\left(p - \\frac12 \\right) \\right) \\\\\n& - a \\theta \\left( \\frac23 + 8 \\left(q - \\frac12 \\right)^2 - \\frac{32}{3} \\left(q - \\frac12 \\right) \\left(p - \\frac12 \\right) \\right) - \\frac89 \\epsilon \\theta \\left( \\left(q - \\frac12 \\right) + \\left(p - \\frac12 \\right) \\right) \n\\end{align*}\\end{small} \nup to a higher order error, when $a$ and $\\epsilon$ are sufficiently small.\nWhile this may not look so useful, we get a great deal of information by looking at the system when $a = \\epsilon = 0$. In such a case, the equations for a critical point in $p,q,\\theta$ become\n\\begin{align*}\n48 \\left(q - \\frac12 \\right) - 16 \\left(p - \\frac12 \\right) + \\frac43 \\theta & = 0 \\\\\n-16 \\left(q - \\frac12 \\right) + 48 \\left(p - \\frac12 \\right) + \\frac43 \\theta & = 0 \\\\\n\\frac43 \\left(q - \\frac12 \\right) + \\frac43 \\left(p - \\frac12 \\right) + \\frac79 \\theta & = 0.\n\\end{align*}\nThus, the Jacobian matrix is\n\\begin{equation*}\nJ_{a=0,\\epsilon=0} = \\left[ \\begin{array}{rrr}\n48 & -16 & \\frac43 \\\\\n-16 & 48 & \\frac43 \\\\\n\\frac43 & \\frac43 & \\frac79\n\\end{array}\n\\right],\n\\end{equation*}\nwhich is non-singular. As a result, we observe that if $a = \\epsilon = 0$, the optimal solutions is $p = q = \\frac12$, $\\theta = 0$. We know this from symmetry arguments, but now we also have set ourselves up for an application of the Implicit Function Theorem in order approximate the RC for near rectangular domains. \n\nWe next observe what happens if $a \\neq 0$, $\\epsilon = 0$. This gives the modified system\n\\begin{equation*}\nJ_{a,\\epsilon=0} \\left[ \\begin{array}{c}\n\\left(q - \\frac12 \\right) \\\\\n\\left(p - \\frac12 \\right) \\\\\n\\theta \n\\end{array} \\right] = a \\left[ \\begin{array}{r}\n8 \\\\\n-8 \\\\\n\\frac23\n\\end{array} \\right] + \\text{Quadratic Error in $a, q-\\frac12, p-\\frac12$},\n\\end{equation*}\nwhere\n\\begin{equation*}\nJ_{a,\\epsilon=0} = \\left[ \\begin{array}{ccc}\n48 -112 a & -16 + 80 a& \\frac43 \\\\\n-16 +80 a & 48 -112 a & \\frac43 \\\\\n\\frac43 & \\frac43 & \\frac79\n\\end{array}\n\\right],\n\\end{equation*}\nHence, $\\vec v$ satisfying $J_{a,\\epsilon=0} \\vec{ v }= \\vec{0}$ is given by \n\\begin{equation*}\n\\vec{v} = a J_{a=0,\\epsilon=0}^{-1} \\left[ \\begin{array}{c}\n8 \\\\ -8 \\\\ \\frac23\n\\end{array} \\right] + O(a^2) = a \\left[ \\begin{array}{c}\n\\frac{1}{12} \\\\ -\\frac16 \\\\ 1 \\end{array} \\right] + O(a^2).\n\\end{equation*}\nNote that the distance between $p$ and $q$ is then $\\frac{a}{4} < a$, with the maximal amplitude of the bulge from the circular arc of size \n\\[ \\frac{a}{8}\\sqrt{y(p)^2+(p-q)^2} < \\frac{a}{8}. \\] \nSince the terms that are linear in $\\epsilon$ are quadratic or higher in $q-1\/2,p-1\/2$, a similar analysis including $\\epsilon$ shows that the curvature of the parabolic curve is actually a lower order deformation for the Ratio Cut than the trapezoidal deflection. Importantly, this argument demonstrates that the trapezoidal deflection is decreasing in the new cut domain. Though the new cut domain will be closer to aspect ratio $1$, the overall deflections are still decreasing on subsequent domains.\n\n\nUsing the Implicit Function Theorem, we are able to compute comparable results for the full Ratio Cut. Indeed, to turn this into a rigorous argument, we need first observe that the minimum Ratio Cut at $a=\\epsilon = 0$ is uniquely $p=q=\\frac12$, $\\theta=0$, which is easily seen by looking at the Hessian. Then, we look at $ \\text{RC} (p,q,\\theta; a, \\epsilon)$ as a map from $\\mathbb{R}^5 \\to \\mathbb{R}^3$, and use the Implicit Function Theorem to construct the desired local map from $(p,q,\\theta;\\,a,\\epsilon)$ in an open set around $(a,\\epsilon) = (0,0)$. \n\n\\subsection{The Ratio Cut with sides given by parabolic approximations}\n\\label{arcs}\n\n\nNow, we proceed to handle a more general family of domains. Given our assumption on the smoothness of the curves, we can assume that the top and bottom curves are approximated by quadratic curves to high accuracy near points of intersection with the ratio cut. Specifically, let us consider an arbitrary domain $Q$ that can be approximated (in the Gromov-Hausdorff sense) by a parabolic trapezoid $Q_0$ with vertices $(0,0),(0,\\frac{1}{2}+a_1),(1,\\frac{1}{2}+a_2),$ and $(1,0)$. Note, the inclusion of $a_1$ and $a_2$ here will allow us to vary the aspect ratio. We fix the width of $Q_0$ to $1$, however, as can always be done by a scaling of the domain. The top paraboloid of $Q_0$ is parametrized as \n$$\ny_{T}(x) = \\epsilon_{t}x^2 + (a_2 - a_1 - \\epsilon_{t})x + a_1 + \\frac{1}{2},\n$$ \nand the bottom paraboloid is parametrized as \n$$\ny_{B}(x) = \\epsilon_{b}x^2 -\\epsilon_b x.\n$$ \nWe assume that $Q$ and $Q_0$ differ by two sufficiently small and bounded ``black-box'' regions on the left and right.\nThe areas of these two small regions will be denoted $A_{\\texttt{WL}}$ and $A_{\\texttt{WR}}$. \n\n\n\\begin{center}\n\\begin{figure}[ht!]\n\\begin{tikzpicture}[scale=4]\n\\draw [ultra thick] (0,1\/2)--(0,0) ;\n\n\\draw [red,ultra thick] (0,0) to[out=5,in=195] (0, 0.5);\n\\node[red] at (-0.15, 0.25) {$A_{\\texttt{WL}}$};\n\n\\draw [ultra thick] (1,0) -- (1, 1\/2+0.1) ;\n\n\\draw [red,ultra thick] (1,0) to[out=105,in=335] (1, 0.6);\n\\node[red] at (1.15, 0.25) {$A_{\\texttt{WR}}$};\n\n\\draw [ultra thick] (0,0.5) to[out=25,in=160] (1, 0.6);\n\\draw [ultra thick] (0,0) to[out=-10,in=190] (1, 0);\n\\filldraw (0,0) circle (0.015cm);\n\\node at (-0.08, -0.08) {(0,0)};\n\\filldraw (1,0) circle (0.015cm);\n\\node at (1.08, -0.08) {(1,0)};\n\\filldraw (1,1\/2+0.1) circle (0.015cm);\n\\node at (1.08, 1\/2+0.1+0.08) {$(1,0.6)$};\n\\filldraw (0,0.5) circle (0.015cm);\n\\node at (-0.1, 0.5+0.08) {($0$,$0.5$)};\n\\filldraw (0.5,-0.05) circle (0.015cm);\n\\node at (0.5, -0.15) {${\\bf q}:=(q,y_B(q))$}; \n\\filldraw (0.52,0.67) circle (0.015cm);\n\\node at (0.48,0.82) {${\\bf p}:=(p,y (p)) $}; \n\\draw [thin,dashed] (0.5,-0.05) to[out=65,in=295] (0.52, 0.67);\n\\draw [dotted] (0.5,-0.05) -- (0.52, 0.67) ;\n\\node at (0.65, 0.48) {$\\Gamma$};\n\\node at (0.55, 0.35) {$\\Omega$};\n\\end{tikzpicture}\n\\caption{A more general domain $Q$ approximated by a parabolic trapezoid.}\n\\label{fig:quad3}\n\\end{figure}\n\\end{center}\n\nWe start by preparing quantities associated to $Q_0$. A circular arc $\\Gamma$ passing through the points $(p,y_{T}(p))$ and $(q, y_B(q))$, with angle $\\theta$, cuts the parabolic trapezoid into a left and right domain, denoted by $S$ and $Q_0\\backslash S$.\n\tTo compute an equivalent analytic expression for the ratio cut, we use Stoke's theorem\n\tto compute the left area $A_L = |S|$ and total $A_T = |Q_0|$, which we use to compute the right area $A_R = |Q_0\\backslash S| = |Q_0|-|S|$. In particular, \n\t$$\n\tA_T = |Q_0| = \\int_{Q_0} dA = \\frac{1}{2} \\int_{\\partial Q_0} xdy - ydx,\n\t$$\n\twhere we integrate along the left vertical boundary $\\{(0,t): 0\\leq t\\leq \\frac{1}{2}+a_1\\}$, along the top parabolic curve $\\{(x,y_T(x)): 0\\leq x\\leq 1 \\},$ along the right vertical boundary $\\{(1,t): \\frac{1}{2}+a_2\\leq t\\leq 0 \\}$ (with the indicated orientation), and finally along the bottom parabolic curve $\\{(x,y_B(x)): 1\\leq x\\leq 0 \\}$ (with the indicated orientation). \n\tWe can compute the (indefinite) integrals for the two parabolic pieces fairly easily:\n\t\\begin{align*}\n\t\tdy_T \t&= (2\\epsilon_t x + a_2-a_1-\\epsilon_t)dx,\\\\\n\t\tdy_B \t&= (2\\epsilon_b x -\\epsilon_b)dx,\n\t\\end{align*}\n\tand so\n\t\\begin{align*}\n\t\t\\frac{1}{2} \\int xdy_T - y_T dx \t&= \\frac{1}{2} \\int x(2\\epsilon_t x + a_2-a_1-\\epsilon_t)dx - (\\epsilon_{t}x^2 + (a_2 - a_1 - \\epsilon_{t})x + a_1 + \\frac{1}{2})dx\\\\\n\t\t\t&= \\frac{1}{2}\\int \\epsilon_t x^2-(a_1+\\frac{1}{2}) dx\n\t\t\t= \\frac{\\epsilon_t}{6}x^3 - \\left(\\frac{a_1}{2}+ \\frac{1}{4} \\right)x;\\\\\n\t\t\\frac{1}{2} \\int xdy_B - y_B dx \t&= \\frac{1}{2}\\int x(2\\epsilon_b x -\\epsilon_b)dx - (\\epsilon_{b}x^2 + -\\epsilon_b x)dx\\\\\n\t\t\t&= \\frac{1}{2} \\int \\epsilon_b x^2 dx\n\t\t\t= \\frac{\\epsilon_b}{6}x^3.\n\t\\end{align*}\nThe $1$-forms for the vertical components are \n\t\\begin{align*}\n\t\t\\frac{1}{2} \\int 0dt - t d(0) &= \\frac{1}{2} \\int 0 = 0,\\\\\n\t\t\\frac{1}{2} \\int 1 dt - t d(1) \t&= \\frac{1}{2} \\int dt - 0 = \\frac{t}{2}.\n\t\\end{align*}\n\tPutting it all together, we have\n\t\\begin{align*}\n\t\tA_T &= 0 + \\left[ \\frac{\\epsilon_t}{6}x^3 - \\left(\\frac{a_1}{2}+ \\frac{1}{4} \\right)x \\right]_{x=0}^{x=\\frac{1}{2}+a_1} + \\left[\\frac{t}{2} \\right]_{t=\\frac{1}{2}+a_2}^{0} + \\left[ \\frac{\\epsilon_b}{6}x^3 \\right]_{x=1}^{x=0}\\\\\n\t\t\t&= \\frac{\\epsilon_t a_1^3}{6} + \\frac{\\epsilon_t a_1^2}{4} + \\frac{\\epsilon_t a_1}{8} + \\frac{\\epsilon_t}{48} -\\frac{\\epsilon_b}{6} - \\frac{a_2}{2} - \\frac{a_1^2}{2} - \\frac{a_1}{2} - \\frac{3}{8}.\n\t\\end{align*}\n\t\n\tFor $A_L = |S|$, we use Stokes' theorem again but split $S$ into two parts divided by the straight line connecting $(p,y_t(p))$ and $(q,y_b(q))$; these parts are a curvilinear quadrilateral and circular cap. The area of the curvilinear quadrilateral is computed by Stokes' theorem, while the area of the circular cap is determined by an elementary formula. Adding these two quantities gives us $A_L$.\n\t%\n\t\n\t\n\tThe line connecting $(p,y_T(p))$ to $(q,y_B(q))$ is parametrized as $\\{ (1-t)(p,y_T(p)) + t(q,y_B(q)): 0\\leq t \\leq 1 \\},$ which componentwise becomes\n\t$$ \n\t( (1-t)p + t q, (1-t)y_T(p) + ty_B(q)), \n\t$$\n\tand so the integral along this straight line becomes\n\t\\begin{align*}\n\t\t\\frac{1}{2} \\int xdy - ydx \t=\\,& \\frac{1}{2} \\int ((1-t)p + t q)d((1-t)y_T(p) + ty_B(q))\\\\ &\\quad - ((1-t)y_T(p) + ty_B(q))d((1-t)p + t q)\\\\\n\t\t\t=\\,& \\frac{1}{2} \\int ((1-t)p + t q)(-y_T(p) dt + y_B(q) dt)\\\\\n\t\t\t\t&\\quad - ((1-t)y_T(p) + ty_B(q)) ( -p dt + q dt)\\\\\n\t\t\t=\\,& t q (p (\\epsilon_b q-\\epsilon_b-\\epsilon_t p+\\epsilon_t)-a_2 (p+1))\\,,\n\t\\end{align*}\n\twhere we denote the last quantity by $SL(t)$.\n\t\n\tGiven two points $p$ and $q$, and an angle $\\theta$, the radius of the circle containing points $p$ and $q$ separated by angle $\\theta$ is \n\t$$\n\tR = \\frac{\\|p-q\\|_2}{2\\sin(\\frac{\\theta}{2})}.\n\t$$ \n\tThus, the area of the circular segment is\t{\\allowdisplaybreaks\n\t$$\n\t|\\Omega|= \\frac{R^2}{2}(\\theta - \\sin(\\theta)) = \\frac{(p - q)^2 + (y_T(p) - y_B(q))^2}{2}(\\theta - \\sin(\\theta)),\n\t$$ \n\tand so\n\t\\begin{align*}\n\t\tA_L =\\,& 0 + \\left[ \\frac{\\epsilon_t}{6}x^3 - \\left(\\frac{a_1}{2}+ \\frac{1}{4} \\right)x \\right]_{x=0}^{x=p} + [SL(t)]_{t=0}^{t=1} + \\left[ \\frac{\\epsilon_b}{6}x^3 \\right]_{x=q}^{x=0} + |\\Omega|\\\\\n\t\t=\\,& \\frac{1}{6} \\left( -\\epsilon_b q^3 - \\frac{3}{2}(1+2a_1)p + \\epsilon_t p^3 + 6q (-a_2(1+p) + p( - \\epsilon_b + \\epsilon_t + \\epsilon_t q - \\epsilon_t p))\\right.\\\\\n\t\t& \\left.+ 3\\left( (q-p)^2 + \\left( a_2 + \\epsilon_b q - \\epsilon_b q^2 + a_2p - \\epsilon_t p + \\epsilon_t p^2\\right)^2\\right) (\\theta - \\sin(\\theta))\\right).\n\t\\end{align*}\n\tWe can use this to get the area of the right portion:\n\t\\begin{align*}\n\t\tA_R =& A_T - A_L \\\\\n\t\t\t=& \\frac{1}{48} \\bigg(-18-24 a_1-24 a_1^2-24 a_2-8 \\epsB+\\epsT+6 a_1 \\epsT+12 a_1^2 \\epsT\\\\\n\t\t\t&+8 a_1^3 \\epsT+8 \\epsB \\tBott^3+12 (1+2 a_1) \\tTop-8 \\epsT \\tTop^3\\\\\n\t\t\t&-48 \\tBott \\big(-a_2 (1+\\tTop)+\\tTop (\\epsT+\\epsB (-1+\\tBott)-\\epsT \\tTop)\\big)\\\\\n\t\t\t&-24 \\big((\\tBott-\\tTop)^2+(a_2+\\epsB \\tBott-\\epsB \\tBott^2+a_2 \\tTop-\\epsT \\tTop+\\epsT \\tTop^2)^2\\big) (\\theta-\\sin(\\theta))\\bigg).\n\t\\end{align*}\n\t}\n\tFinally, the length of the ratio cut takes the form $$|\\Gamma| = R\\theta = \\frac{\\|(p-q, y_T(p)-y_B(q))\\|_2 \\frac{\\theta}{2}}{\\sin(\\frac{\\theta}{2})} = \\frac{\\|(p-q, y_T(p)-y_B(q))\\|_2}{\\sinc(\\frac{\\theta}{2})}.$$\n\n\tWith the above pieces from $Q_0$, the ratio cut of $Q$\n\n\tcan be expressed in terms of variables $p, q,$ and $\\theta$, together with parameters $a_1,a_2, \\epsT, \\epsB, A_{\\texttt{WL}}, A_{\\texttt{WR}}$ that determine the domain. Specifically, by letting $\\sigma = (a_1,a_2, \\epsT, \\epsB, A_{\\texttt{WL}}, A_{\\texttt{WR}})$ be the parameter vector, we can express the ratio cut as \n\t\\begin{align*}\n\t\tRC(q, p, \\theta; \\sigma) \t&= \\frac{\\|(p-q, y_T(p)-y_B(q))\\|_2}{\\sinc(\\frac{\\theta}{2})(A_L + A_{\\texttt{WL}})(A_R+A_{\\texttt{WR}})}.\n\t\\end{align*}\n\t%\n\tThe series expansion of $RC(q, p, \\theta;\\sigma)$ near $\\left( \\frac{1}{2}, \\frac{1}{2}, 0; \\vec{0} \\right)$ is\n\t\\begin{align*}\n\t\tRC \t=& 8 + 24 \\tBCent^2 - 16 \\tBCent \\tTCent + \\frac{4}{3} \\tBCent \\Th\\\\\n\t\t\t&+ 24 \\tTCent^2 + \\frac{4}{3} \\tTCent \\Th + \\frac{7}{18}\\Th^2\\\\\n\t\t\t&+ a_1 p_{a_1}(q, p, \\theta) + a_2 p_{a_2}(q, p, \\theta) + a_3 p_{a_3}(q, p, \\theta)\\\\\n\t\t\t&+ \\epsilon_t p_{\\epsilon_t}(q, p, \\theta) + \\epsilon_b p_{\\epsilon_b}(q, p, \\theta)\\\\\n\t\t\t&+ A_{\\texttt{WL}} p_{A_{\\texttt{WL}}}(q, p, \\theta) + A_{\\texttt{WR}} p_{A_{\\texttt{WR}}}(q, p, \\theta)\\\\\n\t\t\t&+ a_1^{\\,2} p_{a_1 a_1}(q, p, \\theta) + a_1a_2 p_{a_1 a_2}(q, p, \\theta) + a_1a_3 p_{a_1 a_3}(q, p, \\theta)\\\\\n\t\t\t&+ a_2^{\\,2} p_{a_2a_2}(q, p, \\theta) + a_2a_3 p_{a_2a_3}(q, p, \\theta) + a_3^{\\, 2} p_{a_3a_3}(p, q, \\theta)\\\\\n\t\t\t&+ a_1A_{\\texttt{WL}} p_{a_1A_{\\texttt{WL}}}(q, p, \\theta) + a_1A_{\\texttt{WR}} p_{a_1A_{\\texttt{WR}}}(q, p, \\theta)\\\\\n\t\t\t&+ a_1 \\epsT p_{a_1 \\epsT}(q, p, \\theta) + a_1 \\epsB p_{a_1 \\epsB}(q, p, \\theta)\\\\\n\t\t\t&+ a_2 A_{\\texttt{WL}} p_{a_2A_{\\texttt{WL}}}(q, p, \\theta) + a_2A_{\\texttt{WR}} p_{a_2 A_{\\texttt{WR}}}(q, p, \\theta)\\\\\n\t\t\t&+ a_2 \\epsT p_{a_2 \\epsT}(q, p, \\theta) + a_2 \\epsB p_{a_2 \\epsB}(q, p, \\theta)\\\\\n\t\t\t&+ a_3 A_{\\texttt{WL}} p_{a_3A_{\\texttt{WL}}}(q, p, \\theta) + a_3A_{\\texttt{WR}} p_{a_3 A_{\\texttt{WR}}}(q, p, \\theta)\\\\\n\t\t\t&+ a_3 \\epsT p_{a_3 \\epsT}(q, p, \\theta) + a_3 \\epsB p_{a_3 \\epsB}(q, p, \\theta)\\\\\n\t\t\t&+ A_{\\texttt{WL}}A_{\\texttt{WL}} p_{A_{\\texttt{WL}} A_{\\texttt{WL}}}(q, p, \\theta) + A_{\\texttt{WL}} A_{\\texttt{WR}} p_{A_{\\texttt{WL}} A_{\\texttt{WR}}}(q, p, \\theta)\\\\\n\t\t\t&+ A_{\\texttt{WR}} A_{\\texttt{WR}} p_{A_{\\texttt{WR}} A_{\\texttt{WR}}}(q, p, \\theta)\\\\\n\t\t\t&+ A_{\\texttt{WL}} \\epsT p_{A_{\\texttt{WL}} \\epsT}(q, p, \\theta) + A_{\\texttt{WL}} \\epsB p_{A_{\\texttt{WL}} \\epsB}(q, p, \\theta)\\\\\n\t\t\t&+ A_{\\texttt{WR}} \\epsT p_{A_{\\texttt{WR}} \\epsT}(q, p, \\theta) + A_{\\texttt{WR}} \\epsB p_{A_{\\texttt{WR}} \\epsB}(q, p, \\theta)\\\\\n\t\t\t&+ \\epsT \\epsT p_{\\epsT \\epsT}(q, p, \\theta) + \\epsT \\epsB p_{\\epsT \\epsB}(q, p, \\theta) + \\epsB \\epsB p_{\\epsB \\epsB}(q, p, \\theta)\\,,\n\t\\end{align*}\nup to a higher order error, where the polynomial $p_{\\sigma^{\\alpha}}$ is the partial derivative $\\left. \\frac{\\partial^{\\alpha} RC}{\\partial \\sigma^{\\alpha}} \\right|_{\\sigma = 0}$ for a multi-index $\\alpha$. Detailed fomulae for these terms are given in the Appendix.\t\n\t\n\t\n\tNext, we compute the linearization of $RC$ near the point $(q,p,\\theta) = (\\frac{1}{2}, \\frac{1}{2}, 0)$:\n\t\\begin{align*}\n\t\t\\left.\\frac{\\partial RC}{\\partial q}\\right|_{\\sigma = 0} &= 48\\big (q - \\frac{1}{2}\\big) - 16 \\big(p - \\frac{1}{2}\\big) + \\frac{4}{3}\\theta,\\\\\n\t\t\\left.\\frac{\\partial RC}{\\partial p}\\right|_{\\sigma = 0} &= -16 \\big(q - \\frac{1}{2}\\big) +48 \\big(p - \\frac{1}{2}\\big) + \\frac{4}{3}\\theta,\\\\\n\t\t\\left.\\frac{\\partial RC}{\\partial \\theta}\\right|_{\\sigma = 0} &= \\frac{4}{3} \\big(q - \\frac{1}{2}\\big) + \\frac{4}{3} \\big(p - \\frac{1}{2}\\big) + \\frac{7}{9}\\theta.\n\t\\end{align*}\n\t%\n\tThe Jacobian of $RC$ at $\\sigma = 0$ is thus\n\t$$\n\tJ = \\begin{pmatrix} 48 & -16 & \\frac{4}{3}\\\\ -16 & 48 & \\frac{4}{3} \\\\ \\frac{4}{3} & \\frac{4}{3} & \\frac{7}{9} \\end{pmatrix}.\n\t$$\n\t%\nHere, and below, we have kept the $0$-coefficient terms solely as a place keeper to demonstrate that we have actually computed the coefficients of all the terms in our expansion, as well as to make it easier to verify the formulae for the interested reader.\n\tNext, we explore the other pieces of the linearization of $RC$, namely all first-order terms in the variables, together with the parameters.\n\t{\\allowdisplaybreaks\n\tExplicitly, we have\n\t\\begin{align*}\n\n\t\t\\frac{\\partial RC}{\\partial q} \t&= \\left( 48 - 112 a_1 - 112 a_2 + 112 a_3 - 256 A_{\\texttt{WL}} - 256 A_{\\texttt{WR}} - \\frac{200}{3} \\epsilon_b + \\frac{104}{3} \\epsilon_t \\right)\\left( q - \\frac{1}{2} \\right)\\\\\n\t\t&\\quad + \\left(-16 + 80 a_1 + 80a_2 - 80a_3 + 40 \\epsilon_b - 40\\epsilon_t \\right)\\left( p - \\frac{1}{2} \\right)\\\\\n\t\t&\\quad + \\left(\\frac{4}{3} - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t \\right)\\theta\\\\\n\t\t&\\quad + \\Big(8a_1 -8a_2 - 8a_3 + 32 A_{\\texttt{WL}} - 32 A_{\\texttt{WR}} - 32 a_1 a_1 + 0 a_1a_2 + 32 a_1a_3\\\\\n\t\t&\\quad\\quad +32 a_2a_2 + 0 a_2a_3 - 32 a_3a_3 - 128a_1 A_{\\texttt{WL}} + 0 a_1 A_{\\texttt{WR}} + \\frac{8}{3} a_1 \\epsilon_t - \\frac{8}{3} a_1\\epsilon_b\\\\\n\t\t&\\quad \\quad + 0a_2A_{\\texttt{WL}} + 128 a_2 A_{\\texttt{WR}} - \\frac{8}{3} a_2 \\epsilon_t + \\frac{8}{3} a_2 \\epsilon_b\\\\\n\t\t&\\quad \\quad + 64 a_3 A_{\\texttt{WL}} - 64 a_3 A_{\\texttt{WR}} - 8 a_3 \\epsilon_t + 8 a_3 \\epsilon_b\\\\\n\t\t&\\quad \\quad - 512 A_{\\texttt{WL}} A_{\\texttt{WL}} + 0 A_{\\texttt{WL}}A_{\\texttt{WR}} + 512 A_{\\texttt{WR}}A_{\\texttt{WR}} \\\\\n\t\t&\\quad \\quad + \\frac{32}{3} A_{\\texttt{WL}} \\epsilon_t - \\frac{32}{3} A_{\\texttt{WL}} \\epsilon_b - \\frac{32}{3} A_{\\texttt{WR}} \\epsilon_t + \\frac{32}{3} A_{\\texttt{WR}}\\epsilon_b + 0 \\epsilon_t \\epsilon_t + 0 \\epsilon_t \\epsilon_b + 0 \\epsilon_b \\epsilon_b \\Big)\\\\\n\t\t&\\quad + \\text{ cubic terms},\n\t\\end{align*}\n\t\\begin{align*}\n\n\t\t\\frac{\\partial RC}{\\partial p} \t&= \\left(-16 + 80 a_1 + 80a_2 - 80a_3 + 40 \\epsilon_b - 40\\epsilon_t \\right)\\left( q - \\frac{1}{2} \\right)\\\\\n\t\t&\\quad + \\left(48 - 112 a_1 - 112 a_2 +112 a_3 - 256 A_{\\texttt{WL}} - 256 A_{\\texttt{WR}} - \\frac{104}{3} \\epsilon_b + \\frac{200}{3} \\epsilon_t \\right)\\left( p - \\frac{1}{2} \\right)\\\\\n\t\t&\\quad + \\left(\\frac{4}{3} - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t \\right)\\theta\\\\\n\t\t&\\quad + \\left(-8a_1 + 8a_2 + 8a_3 + 32A_{\\texttt{WL}} - 32 A_{\\texttt{WR}} +32 a_1 a_1 + 0 a_1a_2 - 32 a_1a_3\\right.\\\\\n\t\t&\\quad\\quad \\left.-32 a_2a_2 + 0 a_2a_3 + 32 a_3a_3 - 64 a_1 A_{\\texttt{WL}} + 64 a_1 A_{\\texttt{WR}} - 8 a_1 \\epsilon_t + 8 a_1\\epsilon_b\\right.\\\\\n\t\t&\\quad \\quad \\left. - 64 a_2A_{\\texttt{WL}} + 64 a_2 A_{\\texttt{WR}} + 8 a_2 \\epsilon_t - 8 a_2 \\epsilon_b\\right.\\\\\n\t\t&\\quad \\quad \\left. + 0 a_3 A_{\\texttt{WL}} - 128 a_3 A_{\\texttt{WR}} + \\frac{8}{3} a_3 \\epsilon_t - \\frac{8}{3} a_3 \\epsilon_b\\right.\\\\\n\t\t&\\quad \\quad \\left. - 512 A_{\\texttt{WL}} A_{\\texttt{WL}} + 0 A_{\\texttt{WL}}A_{\\texttt{WR}} + 512 A_{\\texttt{WR}}A_{\\texttt{WR}} \\right.\\\\\n\t\t&\\quad \\quad \\left. + \\frac{32}{3} A_{\\texttt{WL}} \\epsilon_t - \\frac{32}{3} A_{\\texttt{WL}} \\epsilon_b + \\frac{32}{3} A_{\\texttt{WR}} \\epsilon_t + \\frac{32}{3} A_{\\texttt{WR}}\\epsilon_b + 0 \\epsilon_t \\epsilon_t + 0 \\epsilon_t \\epsilon_b + 0 \\epsilon_b \\epsilon_b \\right)\\\\\n\t\t&\\quad + \\text{ cubic terms},\n\t\\end{align*}\n\tand\n\t\\begin{align*}\n\n\t\t\\frac{\\partial RC}{\\partial \\theta} \t&= \\left(\\frac{4}{3} - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t \\right)\\left( q - \\frac{1}{2} \\right)\\\\\n\t\t&\\quad + \\left(\\frac{4}{3} - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t \\right)\\left( p - \\frac{1}{2} \\right)\\\\\n\t\t&\\quad + \\left(\\frac{7}{9} - \\frac{5}{9} a_1 - \\frac{5}{9}a_2 + \\frac{5}{9} a_3 - \\frac{32}{9}A_{\\texttt{WL}} - \\frac{32}{9}A_{\\texttt{WR}} + \\frac{1}{54}\\epsilon_b - \\frac{1}{54} \\epsilon_t \\right)\\theta\\\\\n\t\t&\\quad + \\left(\\frac{2}{3}a_1 - \\frac{2}{3}a_2 + \\frac{2}{3} a_3 + \\frac{8}{3}A_{\\texttt{WL}} - \\frac{8}{3}A_{\\texttt{WR}} - \\frac{4}{3} a_1 a_1 + 0 a_1a_2 + 0 a_1a_3\\right.\\\\\n\t\t&\\quad\\quad \\left.+ \\frac{4}{3} a_2a_2 - \\frac{4}{3} a_2a_3 + \\frac{4}{3} a_3a_3 - 8 a_1 A_{\\texttt{WL}} - \\frac{8}{3} a_1 A_{\\texttt{WR}} - \\frac{1}{9} a_1 \\epsilon_t + \\frac{1}{9} a_1\\epsilon_b\\right.\\\\\n\t\t&\\quad \\quad \\left.+ \\frac{8}{3} a_2A_{\\texttt{WL}} + 8 a_2 A_{\\texttt{WR}} + \\frac{1}{9} a_2 \\epsilon_t - \\frac{1}{9} a_2 \\epsilon_b\\right.\\\\\n\t\t&\\quad \\quad \\left. - \\frac{8}{3} a_3 A_{\\texttt{WL}} - 8 a_3 A_{\\texttt{WR}} - \\frac{1}{9} a_3 \\epsilon_t + \\frac{1}{9} a_3 \\epsilon_b\\right.\\\\\n\t\t&\\quad \\quad \\left. - \\frac{128}{3} A_{\\texttt{WL}} A_{\\texttt{WL}} + 0 A_{\\texttt{WL}}A_{\\texttt{WR}} + \\frac{128}{3} A_{\\texttt{WR}}A_{\\texttt{WR}} \\right.\\\\\n\t\t&\\quad \\quad \\left. - \\frac{4}{9} A_{\\texttt{WL}} \\epsilon_t + \\frac{4}{9} A_{\\texttt{WL}} \\epsilon_b + \\frac{4}{9} A_{\\texttt{WR}} \\epsilon_t - \\frac{4}{9} A_{\\texttt{WR}}\\epsilon_b + 0 \\epsilon_t \\epsilon_t + 0 \\epsilon_t \\epsilon_b + 0 \\epsilon_b \\epsilon_b \\right)\\\\\n\t\t&\\quad + \\text{ cubic terms}.\n\t\\end{align*}\n\t%\n\tWith these expansions, we can express the linearization as $J+ J_{\\sigma}$, where the terms $J_{\\sigma,ii}$ come from the partials computed above. Explicitly,\n\t\\begin{align*}\n\t\tJ_{\\sigma,11} \t&= - 112 a_1 - 112 a_2 - 256 A_{\\texttt{WL}} - 256 A_{\\texttt{WR}} - \\frac{200}{3} \\epsilon_b + \\frac{104}{3} \\epsilon_t,\\\\\n\t\tJ_{\\sigma,12}\t&= 80 a_1 + 80a_2 + 40 \\epsilon_b - 40\\epsilon_t,\\\\\n\t\tJ_{\\sigma,13} \t&= - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t,\\\\\n\t\tJ_{\\sigma,21} \t&= 80 a_1 + 80a_2 + 40 \\epsilon_b - 40\\epsilon_t,\\\\\n\t\tJ_{\\sigma,22} \t&= - 112 a_1 - 112 a_2 - 256 A_{\\texttt{WL}} - 256 A_{\\texttt{WR}} - \\frac{104}{3} \\epsilon_b + \\frac{200}{3} \\epsilon_t,\\\\\n\t\tJ_{\\sigma,23} \t&= - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t,\\\\\n\t\tJ_{\\sigma,31} \t&= - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t,\\\\\n\t\tJ_{\\sigma,32} \t&= - \\frac{32}{3}A_{\\texttt{WL}} - \\frac{32}{3}A_{\\texttt{WR}} + \\frac{8}{9} \\epsilon_b - \\frac{8}{9}\\epsilon_t,\\\\\n\t\tJ_{\\sigma,33} \t&= - \\frac{5}{9} a_1 - \\frac{5}{9}a_2 - \\frac{32}{9}A_{\\texttt{WL}} - \\frac{32}{9}A_{\\texttt{WR}} + \\frac{1}{54}\\epsilon_b - \\frac{1}{54} \\epsilon_t.\n\t\\end{align*}\n\t\n\t\n\tWith these pieces we can express the criterion for being a critical point of the Ratio Cut, incorporating linear terms in the parameters, as\n\t\\begin{align*}\n\t\t(J+J_{\\sigma}) \\begin{pmatrix} q - \\frac{1}{2}\\\\ p - \\frac{1}{2}\\\\ \\theta \\end{pmatrix} &= L(\\sigma)\\\\\n\t\t&:= a_1 \\begin{pmatrix} 8 \\\\ -8 \\\\ \\frac{2}{3}\\end{pmatrix} + a_2 \\begin{pmatrix}-8\\\\ 8\\\\ -\\frac{2}{3}\\end{pmatrix} + a_3 \\pmat{-8 \\\\ 8 \\\\ \\frac{2}{3}} + A_{\\texttt{WL}} \\begin{pmatrix}32 \\\\ 32 \\\\ \\frac{8}{3} \\end{pmatrix} + A_{\\texttt{WR}} \\begin{pmatrix}-32\\\\ -32 \\\\ -\\frac{8}{3}\\end{pmatrix}\\\\\n\t\t&\\quad + a_1a_1 \\pmat{-32 \\\\ 32 \\\\ - \\frac{4}{3}} + a_1a_2 \\pmat{0 \\\\ 0 \\\\ 0} + a_1a_3 \\pmat{32 \\\\ -32 \\\\ 0} \\\\\n\t\t&\\quad + a_2a_2 \\pmat{32 \\\\ -32 \\\\ \\frac{4}{3}} + a_2a_3 \\pmat{0 \\\\ 0 \\\\ -\\frac{4}{3}} + a_3a_3 \\pmat{ - 32 \\\\ 32 \\\\ \\frac{4}{3} } \\\\\n\t\t&\\quad \t+ a_1A_{\\texttt{WL}} \\pmat{ -128 \\\\ -64 \\\\ -8 } + a_1 A_{\\texttt{WR}} \\pmat{ 0 \\\\ 64 \\\\ - \\frac{8}{3} } + a_1 \\epsilon_t \\pmat{ \\frac{8}{3} \\\\ -8 \\\\ - \\frac{1}{9} } + a_1 \\epsilon_b \\pmat{ - \\frac{8}{3} \\\\ 8 \\\\ \\frac{1}{9} } \\\\\n\t\t&\\quad \t+ a_2 A_{\\texttt{WL}} \\pmat{ 0 \\\\ -64 \\\\ \\frac{8}{3} } + a_2 A_{\\texttt{WR}} \\pmat{ 128 \\\\ 64 \\\\ 8 } + a_2 \\epsilon_t \\pmat{ - \\frac{8}{3} \\\\ 8 \\\\ \\frac{1}{9} } + a_2 \\epsilon_b \\pmat{ \\frac{8}{3} \\\\ - 8 \\\\ -\\frac{1}{9} } \\\\\n\t\t&\\quad + a_3 A_{\\texttt{WL}} \\pmat{ 64 \\\\ 0 \\\\ -\\frac{8}{3} } + a_3 A_{\\texttt{WR}} \\pmat{ -64 \\\\ - 128 \\\\ -8 } + a_3 \\epsilon_t \\pmat{ -8 \\\\ \\frac{8}{3} \\\\ -\\frac{1}{9} } + a_3 \\epsilon_b \\pmat{ 8 \\\\ -\\frac{8}{3} \\\\ \\frac{1}{9} } \\\\\n\t\t&\\quad + A_{\\texttt{WL}} A_{\\texttt{WL}} \\pmat{ -512 \\\\ -512 \\\\ -\\frac{128}{3} } + A_{\\texttt{WL}} A_{\\texttt{WR}} \\pmat{ 0 \\\\ 0 \\\\ 0 } + A_{\\texttt{WR}} A_{\\texttt{WR}} \\pmat{ 512 \\\\ 512 \\\\ \\frac{128}{3} } \\\\\n\t\t&\\quad + A_{\\texttt{WL}} \\epsilon_t \\pmat{\\frac{32}{3} \\\\ \\frac{32}{3} \\\\ - \\frac{4}{9} } + A_{\\texttt{WL}} \\epsilon_b \\pmat{ - \\frac{32}{3} \\\\ - \\frac{32}{3} \\\\ \\frac{4}{9} } \\\\\n\t\t&\\quad + A_{\\texttt{WR}} \\epsilon_t \\pmat{ - \\frac{32}{3} \\\\ \\frac{32}{3} \\\\ \\frac{4}{9} } + A_{\\texttt{WR}} \\epsilon_b \\pmat{ \\frac{32}{3} \\\\ \\frac{32}{3} \\\\ -\\frac{4}{9} } \\\\\n\t\t&\\quad + \\epsilon_t \\epsilon_t \\pmat{0\\\\0\\\\0 } + \\epsilon_t \\epsilon_b \\pmat{0\\\\ 0\\\\ 0 } + \\epsilon_b \\epsilon_b \\pmat{0\\\\ 0 \\\\ 0 }\\\\\n\t\t&\\quad + \\text{ cubic terms in the parameters}.\n\t\\end{align*}\n\t}\n\t\\subsection*{Using the Implicit Function Theorem}\n\t\n\tSuppose $v$ is a critical triple of values for the Ratio Cut, i.e. $(J+J_{\\sigma})v = 0$. Then near $\\sigma = 0$, we can solve for $v$ in terms of the parameters in $\\sigma$:\n\t\\begin{align*}\n\t\tv \t&= J^{-1} L(\\sigma)\\\\\n\t\t\t&= \\pmat{\\frac{1}{12}(a_1 - a_2 - 2a_3) + (A_{\\texttt{WL}} - A_{\\texttt{WR}}) \\\\ -\\frac{1}{12} (2a_1 - 2a_2 - a_3) + (A_{\\texttt{WL}} - A_{\\texttt{WR}}) \\\\ a_1 - a_2 + a_3} + \\text{quadratic terms in the parameters}.\n\t\\end{align*}\n\t%\n\tWe will not write out the full $J^{-1}L(\\sigma)$ here, though it is used for the approximations later and its form becomes clear in the representations below.\n\t%\n\t\n\t\n\tTo summarize, we have shown in this section that, near the critical point $(p,q,\\theta) = \\left( \\frac{1}{2}, \\frac{1}{2}, 0 \\right)$ of the ratio cut, and for $\\sigma=(a_1, a_2, \\epsilon_b, \\epsilon_t, A_{\\texttt{WL}}, A_{\\texttt{WR}})$ near $0$, we have \n\t\\begin{align*}\n\t\t\\begin{pmatrix} q - \\frac{1}{2} \\\\ p - \\frac{1}{2} \\\\ \\theta \\end{pmatrix} = \\pmat{\\frac{1}{12}(a_1 - a_2 - 2a_3) + (A_{\\texttt{WL}} - A_{\\texttt{WR}}) \\\\ -\\frac{1}{12} (2a_1 - 2a_2 - a_3) + (A_{\\texttt{WL}} - A_{\\texttt{WR}}) \\\\ a_1 - a_2 + a_3}+ \\text{quadratic terms in the parameters}.\n\t\\end{align*}\n\t\n\t\t\n\t\\section{Are our domains generic?}\n\t\n\tSince Ratio Cuts are circular, it seems more natural to prove Theorem 2 for circular, instead of parabolic, boundary curves. Using circular arcs, however, proved to be intractable for our computational methods, whereas parabolic arcs could be used. This section shows that such approximations only incur third order errors, and thus Theorem 2 is still applicable for circular boundary curves. This first lemma shows this approximation is valid near the intersection of the cut and the original domain's boundary. \n\t\n\t\\begin{lemma}\n\tProvided $I(Q)$ defined in \\eqref{Definition I(Q)} is sufficiently small, we may approximate the top and bottom curves by parabolas in a neighborhood of the Ratio Cut.\n\t\\end{lemma}\n\t\n\t\\begin{proof}\n\tThe small $I(Q)$ makes sure the domain is approximately trapezoidal with a, for example, $1:2$, aspect ratio, and that the top and curves are $C^3$ with small oscillations on a uniform spatial scale. Hence, Taylor expanding the top and bottom curves out to $2$nd order on this scale near $p=1\/2, q = 1\/2$, we can create a parabolic approximation that is accurate up to $O( |p-\\frac12|^3, |q-\\frac12|^3)$ in this region. Potentially modifying the left and right wings in order to ensure the wings of domain have the appropriate error outside this uniform neighborhood, we see that these curvilinear trapezoids will have ratio cuts that are indeed approximated up to lower order terms in the same fashion as our exact parabolic trapezoid calculations. \n\t\\end{proof}\n\t\t\n\tIn practice, we are given a domain $Q$ whose top and bottom boundary curves are more general to start, and upon iteration of our domains we are most interested when sides are given by circular arcs. To handle smooth enough domains of this form in full generality, we can construct a parabolic trapezoid by: rescaling the domain to fit our aspect ratio; locally using a parabolic approximation for the top and bottom curves; and finally cutting off left and right portions of $\\Omega$, which are treated as black-box regions. The Ratio Cut function sees these regions as general wing areas, denoted $A_{\\texttt{WL}}$ and $A_{\\texttt{WR}}$ for the left- and right-wing areas respectively. The measure of distance from the rectangle, $I(Q)$, in which we measure the deflection of our domains, allows us to approximate our quadrilateral by a parabolic curve with errors that are higher order in the parameter space. \n\t\t\n\t\\subsection{Can a circular arc be parabolically approximated?}\n\t\n\tOptimal cuts for generic parabolic trapezoids are circular. Because of this, we would expect the top and bottom boundary curves of the domains in the next iteration we have been considering to be circular, but not parabolic. Here we show that parabolic arcs approximate circular arcs well, and that the ratio cut for circular boundary curves differs from the ratio cut for parabolic boundary curves at third order and higher. Thus, working with parabolic curves does not incur any significant error in the ratio cut series expansion.\n\t\n\t\\begin{lemma}\n\t Up to higher order error terms in the curve parameters, we may approximate the top and bottom circular arc curves as paraboloids encompassing the correct area.\n\t \\end{lemma}\n\t \n\t The remainder of this section will be devoted to the proof of this Lemma.\n\t\n\t\\subsubsection{Approximating Circular Arcs} \n\tFirst let us consider the most important example for our iterated domain conjecture, in which the top and bottom boundaries are circular arcs. In what follows, the top and bottom parabolic curves will be, respectively,\n\t\\begin{align*}\n\t\ty_T^P(x) \t&= \\epsilon_t x^2 + (a_1 - a_2 - \\epsilon_t)x+ \\left( a_1 + \\frac{1}{2} \\right),\\\\\n\t\ty_B^P(x) \t&= \\epsilon_b x^2 + (a_3 - \\epsilon_b)x.\n\t\\end{align*}\n\t\n\tThe parameters $a_1, a_2, a_3$ denote horizontal perturbations from the top-left, top-right, and bottom-right vertices of a rectangle, so our parabolic trapezoid has vertices $(0,0)$, $\\left(0, \\frac{1}{2}+a_1 \\right)$, $\\left(1, \\frac{1}{2}+a_2 \\right),$ and $\\left(1, a_3 \\right)$. The terms $\\epsilon_t$ and $\\epsilon_b$ are curvature parameters that specify the shape of the two boundary curves.\n\t\n\tThe circular arcs are given by the formulas\n\t\\begin{align*}\n\t\ty_T^C(x) &= c_{y,t} + \\sqrt{r_t^2 - (x-c_{x,t})^2}, \\\\\n\t\ty_B^C(x) \t&= c_{y,b} - \\sqrt{r_b^2 - (x-c_{x,b})^2},\n\t\\end{align*}\n\twhere $r_i$ and $(c_{x,i},c_{y,i})$ are the radius and center of the corresponding top ($i=T$) or bottom ($i=B$) circle:\n\t\n\t\\begin{align*}\n\t\tr_t &= \\frac{\\sqrt{1 + (a_1-a_2)^2}}{2 \\sin(\\frac{\\theta}{2})},\\\\\n\t\tr_b &= \\frac{\\sqrt{1+a_3^2}}{2 \\sin(\\frac{\\theta}{2})},\n\t\\end{align*}\n\tand the center point coordinates are found from the systems of equations (for $i=T$ and $i=B$)\n\t\\begin{align*}\n\t\t&\\begin{cases}\n\t\t(c_{x,t})^2 + (\\frac{1}{2} + a_1 - c_{y,t})^2 \t= r_t^2,\\\\\n\t\t(1-c_{x,t})^2 + (\\frac{1}{2}+a_2 - c_{y,t})^2 \t= r_t^2, \n\t\t\\end{cases}\\\\\n\t\t\\text{and } \n\t\t&\\begin{cases}\n\t\t(c_{x,b})^2 + (c_{y,b})^2 \t= r_b^2,\\\\\n\t\t(1-c_{x,b})^2 + (a_3 - c_{y,b})^2 \t= r_b^2.\n\t\t\\end{cases}\n\t\\end{align*}\n\t\n\tAs an explicit example, expressing $y_B^C$ in terms of the parameters gives\n\t\\begin{align*}\n\n\t\ty_B^C(x) \t&= \\frac{a_3}{2} + \\cot\\left( \\frac{\\theta_b}{2} \\right) - \\left( \\frac{1}{1+a_3^2} \\left( a_3^4 - (1-2x)^2 - 4a_3^2 (x-1) x \\right. \\right. \\\\\n\t\t&\\quad \\left.\\left. + 2a_3 (1+a_3^2) (1-2x) \\cot\\left( \\frac{\\theta_b}{2} \\right) + (1+a_3^2) \\csc^2\\left( \\frac{\\theta_b}{2} \\right) \\right) \\right)^{\\frac{1}{2}}.\n\t\\end{align*}\n\t\n\tThe next lemma shows how to choose $\\epsilon_t$ and $\\epsilon_b$ so that the parabolic trapezoid's area approximates the circular trapezoid's area up to third order. These curvature terms are chosen so that the quadrilateral's parabolic caps have, up to third order terms, the same area as corresponding circular arcs.\n\t\n\t\\begin{lemma}\n\tLet $l_T(x) = (1-t)\\left(\\frac{1}{2}+a_1 \\right) + t \\left(\\frac{1}{2}+a_2 \\right)$, and write \n\t\\begin{equation}\n\tS_T^C = \\{ (x,y) \\colon l_T(x) \\leq y \\leq y_T^C(x) \\}, \\,\\, S_T^P(x) = \\{ (x,y) \\colon l_T(x) \\leq y \\leq y_T^P(x) \\}\n\t\\end{equation} \n\tfor the regions bounded below by the straight line $\\{\\{ t, l_T(x) \\} \\colon t \\in [0,1] \\}$ and above by the curves $\\{\\{t, y_T^i(x) \\} \\colon x\\in[0,1], i = C \\text{ or } P \\}$ respectively. Similarly, let $l_B(x) = t a_3,$ and write\n\t \\begin{equation}\n\tS_B^C = \\{ (x,y) \\colon l_B(x) \\leq y \\leq y_B^C(x) \\},\\,\\, S_B^P(x) = \\{ (x,y) \\colon l_B(x) \\leq y \\leq y_B^P(x) \\}\n\t\\end{equation} \n\tfor the regions bounded above by the straight line $\\{\\{ t, l_B(x) \\} \\colon t \\in [0,1] \\}$ and below by the curves $\\{\\{t, y_B^i(x) \\} \\colon x\\in[0,1], i = C \\text{ or } P\\}$ respectively.\n\t\n\t\tFor $\\epsilon_t = -\\frac{1+(a_1-a_2)^2}{2}\\theta_t$ and $\\epsilon_b = \\frac{1+a_3^2}{2}\\theta_b$, we have\n\t\t\\begin{align*}\n\t\t\t|S_T^C| -|S_T^P| \t&= O(\\theta_t^3)\n\t\t\t\\end{align*}\n\t\t\tand\n\t\t\t\\begin{align*}\n\t\t\t\t\t\t|S_B^C| - |S_B^P| \t&= O(\\theta_b^3) .\n\t\t\\end{align*}\n\t\\end{lemma}\n\t\n\t\\begin{proof}\n\t\tBoth $|S_T^C|$ and $|S_B^C|$ can be found using basic trigonometry: the area of a circular segment with radius $r$ and angle $\\theta$ is $\\frac{1}{2} r^2 (\\theta - \\sin(\\theta))$. In our case,\n\t\t\\begin{align*}\n\t\t\t|S_T^C| \t&= \\frac{1}{2} \\left(\\frac{\\sqrt{1+(a_1-a_2)^2}}{2\\sin(\\frac{\\theta_t}{2})} \\right)^2 (\\theta_t - \\sin(\\theta_t) )\\\\\n\t\t\t\t&= \\frac{1+(a_1-a_2)^2}{12}\\theta_t + \\frac{1+(a_1-a_2)^2}{360}\\theta_t^3 + O(\\theta_t^4),\\\\\n\t\t\t|S_B^C| \t&= \\frac{1}{2} \\left(\\frac{\\sqrt{1+a_3^2}}{2\\sin(\\frac{\\theta_b}{2})} \\right)^2(\\theta_b - \\sin(\\theta_b))\\\\\n\t\t\t\t&= \\frac{1+a_3^2}{12} \\theta_b + \\frac{1+a_3^2}{360} \\theta_b^3 + O(\\theta_b^4).\\\\\n\t\t\\end{align*}\n\t\t\n\t\tFor $|S_T^P|$ and $|S_B^P|$, we integrate:\n\t\t\\begin{align*}\n\t\t\t|S_T^P| \t&= \\int_{0}^1 y_T^P(x) - l_T(x)dx\\\\\n\t\t\t\t&= \\int_0^1 \\epsilon_t x^2 + (a_2 - a_1 - \\epsilon_t)x + \\left( \\frac{1}{2} + a_1 \\right) - (1-x)\\left( \\frac{1}{2}+a_1 \\right) - x\\left(\\frac{1}{2}+a_2 \\right) dx\\\\\n\t\t\t\t&= -\\frac{\\epsilon_t}{6}\n\t\t\t\t\\end{align*}\n\t\tand \n\t\t\\begin{align*}\n\t\t |S_B^C| \t&= \\int_0^1 l_B(x) - y_B^C(x) dx\\\\\n\t\t\t&= \\int_0^1 a_3 x -\\epsilon_b x^2 - (a_3 - \\epsilon_b)x dx\n\t\t\t= \\frac{\\epsilon_b}{6}.\n\t\t\\end{align*}\n\t\tWe want to choose constants $C_t$ and $C_b$ so that setting $\\epsilon_t = C_t \\theta_t$ and $\\epsilon_b = C_b \\theta_b$ gives us $|S_T^C|-|S_T^P| = O(\\theta_t^3)$ and $|S_B^C| - |S_B^P| = O(\\theta_b^3)$. Indeed, letting $\\epsilon_t = -\\frac{1+(a_1-a_2)^2}{2}\\theta_t$ and $\\epsilon_b = \\frac{1+a_3^2}{2}\\theta_b$ gives us this result.\n\t\\end{proof}\n\t\n\tSince the parabolic and circular trapezoids only differ in the type of caps on each, making these choices for $\\epsilon_t$ and $\\epsilon_b$ ensures the area of the circular and parabolic trapezoids are very well approximated. In addition, the boundary curves agree up to third order, and so the left areas and ratio cut lengths also agree to at least third order (in the parameters).\n\t\\begin{cor} \n\t\tSetting $\\epsilon_t = -\\frac{1+(a_1-a_2)^2}{2}\\theta_t$ and $\\epsilon_b = \\frac{1+a_3^2}{2}\\theta_b$, we get the following approximations:\n\t\t\\begin{align*}\n\t\ty_T^C(x) - y_T^P(x) &= \\frac{1}{48}(1+(a_1-a_2)^2) \\theta_t^2 \\left( O(a_1) + O(a_2) + O(\\theta_t) \\right),\\\\\n\t\ty_B^C(x) - y_B^P(x) &= \\frac{1}{48} (1+a_3^2) \\theta_b^2 (O(a_3) + O(\\theta_b)).\n\t\t\\end{align*}\n\t\t\n\t\tMoreover, we have\n\t\t\\begin{align*}\n\t\tA_T^C \t&= A_T^P + O(\\theta_t^3) + O(\\theta_b^3),\\\\\n\t\t\\text{and } A_L^C \t&= A_L^P + O(\\theta_t^2)(O(a_1) + O(a_2)) + O( \\theta_b^2)O(a_3),\n\t\t\\end{align*}\n\t\twhere $A_T^i$, $A_L^i$ is the total area and left area respectively of the parabolic ($i=P$) or circular ($i=C$) trapezoid, and\n\t\t\\begin{align*}\n\t\t\t\\|(p,y_T^C(p)) - (q,y_B^C(q)) \\| \t&= \\|(p,y_T^P(p)) - (q,y_B^P(q))\\| + O(\\theta_t^3)+O(\\theta_b^3).\n\t\t\\end{align*}\n\t\\end{cor}\n\t\n\tThe corollary justifies our use of only quadratic terms in the Ratio Cut Taylor expansion, since the parabolic and circular Ratio Cuts (i.e. Ratio Cut functions of parabolic or circular trapezoids) agree up to third order in the parameters of the domains, given the proper curvature term substitutions.\n\t\n\t\\begin{proof}\n\t\tThe parabolic and circular boundary curve approximations come by Taylor expansions.\n\t\t%\n\t\tFor the areas, recall that the parabolic and circular trapezoids differ only in their ``caps''. Thus, assuming $a_1 B_\\star^{\\rm drip}\\equiv \\frac{1}{2}\\left(\\frac{ \\mu_e^{\\rm drip}(A,Z)}{m_e c^2}\\right)^2 \n\\biggl[1-\\frac{8 C \\alpha Z^{2\/3}}{3(2 \\pi^2)^{1\/3}} \\biggr] \\, .\n\\end{equation}\n\n\n\\subsection{Accreting neutron stars}\n\nIn accreting neutron stars, the magnetic field is typically negligibly small ($B\\ll B_\\star$) and will thus be ignored. \nFor an accretion rate $\\dot{M}=10^{-9}~{\\rm M_\\odot}$~yr$^{-1}$ the original outer crust is replaced by accreted matter in $10^4$~yr.\nFor low-mass binary systems, the accretion stage can last for $10^9$~yr.\nAt densities above $\\sim 10^8$~g~cm$^{-3}$, matter is highly degenerate and relatively cold ($T\\lesssim 5\\times 10^8$~K) so that thermonuclear processes are strongly suppressed, since their rates are many orders of magnitude lower than the compression rate due to accretion \\cite{haensel2007}. The only relevant \nprocesses are electron captures and neutron-emission processes, whereby the nucleus $^A_ZX$ is transformed into a nucleus $^{A-\\Delta N}_{Z-1}Y$ with \nproton number $Z-1$ and mass number $A-\\Delta N$ by capturing an electron with the emission of $\\Delta N$ neutrons $n$ and an electron neutrino $\\nu_e$:\n\\begin{equation}\n\\label{eq:e-capture+n-emission}\n^A_ZX+ e^- \\rightarrow ^{A-\\Delta N}_{Z-1}Y+\\Delta N n+ \\nu_e\\, .\n\\end{equation}\nIn this case, the condition for the onset of neutron drip becomes~\\cite{chamel2015a}\n\\begin{eqnarray}\n\\label{eq:e-capture+n-emission-gibbs-approx}\n\\mu_e + C e^2 n_e^{1\/3}\\biggl[Z^{5\/3}-(Z-1)^{5\/3} + \\frac{1}{3} Z^{2\/3}\\biggr] = \\mu_e^{\\rm drip-acc} \\, ,\n\\end{eqnarray}\nwhere\n\\begin{equation}\n\\label{eq:muebetan}\n\\mu_e^{\\rm drip-acc}(A,Z)\\equiv M'(A-\\Delta N,Z-1)c^2-M'(A,Z)c^2 +m_n c^2 \\Delta N + m_e c^2 \\, .\n\\end{equation}\nThe neutron-drip density and pressure are approximately given by~\\cite{chamel2015a}\n\\begin{equation}\n\\label{eq:ndrip-acc}\nn_{\\rm drip-acc}(A,Z) \\approx \\frac{A}{Z} \\frac{\\mu_e^{\\rm drip-acc}(A,Z)^3}{3\\pi^2 (\\hbar c)^3} \n \\biggl[1+\\frac{C \\alpha}{(3\\pi^2)^{1\/3}}\\left(Z^{5\/3}-(Z-1)^{5\/3}+\\frac{Z^{2\/3}}{3}\\right)\\biggr]^{-3}\\, ,\n\\end{equation}\n\\begin{equation}\n\\label{eq:Pdrip-acc}\nP_{\\rm drip-acc}(A,Z) \\approx \\frac{\\mu_e^{\\rm drip-acc}(A,Z)^4}{12 \\pi^2 (\\hbar c)^3}\\biggl[1+\\frac{4C \\alpha Z^{2\/3}}{(81\\pi^2)^{1\/3}} \\biggr]\n \\biggl[1+\\frac{C \\alpha}{(3\\pi^2)^{1\/3}}\\left(Z^{5\/3}-(Z-1)^{5\/3}+\\frac{Z^{2\/3}}{3}\\right)\\biggr]^{-4}\\, .\n\\end{equation}\nAs discussed in Ref.~\\cite{chamel2015a}, the dripping nucleus can be determined as follows. Given the mass number $A$ and the initial atomic number \n$Z_0$ of the ashes of x-ray bursts, the atomic number $Z$ at the neutron-drip point is the highest number of protons lying below $Z_0$ for which the $\\Delta N$-neutron \nseparation energy defined as\n\\begin{equation}\n\\label{eq:sn-dn}\nS_{\\Delta N n}(A,Z-1) \\equiv M(A-\\Delta N,Z-1)c^2 - M(A,Z-1)c^2 + \\Delta N m_n c^2\n\\end{equation}\nis negative. \n\n\n\n\\section{Numerical results}\n\\label{sec:results}\n\n\\subsection{Nonaccreting neutron stars}\n\\label{sec:results-nonaccreting}\n\nWe have calculated the properties of neutron-star crusts at the neutron-drip point by minimizing the Gibbs free energy per nucleon~(\\ref{eq:gibbs}), \nboth in the absence and in the presence of a strong magnetic field. In the latter case, we have set $B_\\star=500$, $1000$, $1500$ and $2000$ corresponding \nto magnetic fields in the range $2.2 \\times 10^{16}$~G to $8.8 \\times 10^{16}$~G. Since nuclear masses at this depth of the outer crust are not experimentally \nknown, the predictions for the dripping nucleus are model dependent. The neutron-drip properties are summarized in Table~\\ref{tab:drip-cat} in \nunmagnetized neutron stars, and in Tables~\\ref{tab:drip-cat-mag-500}-\\ref{tab:drip-cat-mag-2000} in strongly magnetized neutron stars (magnetars). \n\nFigure~\\ref{fig:cat-ndrip-B} shows that for any given value of the magnetic field strength, the neutron-drip density increases almost linearly with the slope \nof the symmetry energy $L$ (or equivalently with $J$ since the two coefficients are strongly correlated, as previously discussed in Sec.~\\ref{sec:hfb-models}). \nOn the other hand, the behavior of the neutron-drip density with respect to the magnetic field strength exhibits typical quantum oscillation whereas the \nneutron-drip pressure increases monotonically, as recently discussed in Ref.~\\cite{chamel2015b}. The errors of the analytical formulas~(\\ref{eq:ndrip-cat-approx})-(\\ref{eq:Pdrip-cat-approx}) amount to $0.1\\%$ at most, as compared to the numerical solution of Eq.~(\\ref{eq:n-drip-mue}). The proton fraction $Z\/A$ \nat the neutron-drip point is also found to be strongly correlated with the symmetry energy. As shown in the right panel of Fig.~\\ref{fig:cat-AZ-B}, $Z\/A$ decreases \nalmost linearly with increasing $L$ (or $J$). Similar behaviors of $Z\/A$ and $n_{\\rm drip}$ with $L$ have been recently obtained in Ref.~\\cite{bao2014}, and \ncan be inferred from the discussions in Refs.~\\cite{roca2008,provid2014,grill2014}. Nevertheless, in all cases they considered the limiting case $B_\\star=0$. \nIn Ref.~\\cite{bao2014}, the authors studied the role of the symmetry energy on the properties of neutron-star crusts around \nthe neutron-drip threshold using two sets of relativistic mean field (RMF) models based on the TM1 and IUFSU parametrizations respectively. They generated \nseries of models so as to achieve different values of $L$ keeping the symmetry energy at $n=0.11$~fm$^{-3}$ fixed. In our case, the fixed value of the symmetry \nenergy at $n\\approx 0.11$~fm$^{-3}$ results from the mass fit without any further constraint. Although the variations of $Z\/A$ and $n_{\\rm drip}$ they found \nare nonlinear over this range of values of $L$, the variations become almost linear on the narrower range we consider (from about 37~MeV to about 69~MeV). \nAlthough it has been found that a soft symmetry energy favors neutron drip in isolated nuclei~\\cite{todd2003}, this result does not necessarily imply the observed \ncorrelation between $n_{\\rm drip}$ and $L$. Indeed, as recently discussed in Ref.~\\cite{chamel2015a}, the dripping nucleus in the crust is actually stable \nagainst neutron emission, but unstable against electron captures followed by neutron emission. Actually, as will be discussed in Sec.~\\ref{sec:results-accreting}, \naccreting neutron star crusts exhibit different correlations between $n_{\\rm drip}$ and $L$. \n\n\n\\begin{figure*}\n\\begin{center}\n\\includegraphics[scale=0.45]{drip_cat_22-25_B_ndrip.eps}\n\\end{center}\n\\caption{(Color online) Neutron-drip density in nonaccreting neutron-star crusts as a function of the slope $L$ of the symmetry energy of infinite homogeneous nuclear matter at saturation\nand for different magnetic field\nstrengths, as obtained using the HFB-22 to HFB-25 Brussels-Montreal nuclear mass models~\\cite{goriely2013}.}\n\\label{fig:cat-ndrip-B}\n\\end{figure*}\n\n\\begin{figure*}\n\\begin{center}\n\\includegraphics[scale=0.45]{drip_cat_22-25_AZ_B.eps}\n\\end{center}\n\\caption{(Color online) (a) Mass number $A$ and (b) proton fraction $Z\/A$ at the neutron-drip transition for nonaccreting neutron-star crusts as a function of the slope $L$ of the \nsymmetry energy of infinite homogeneous nuclear matter at saturation, \nas obtained using the HFB-22 to HFB-25 Brussels-Montreal nuclear mass models~\\cite{goriely2013}. Squares (circles) correspond to $B_\\star=0, 500$ and $1500$ \n($B_\\star =1000$ and $2000$).}\n\\label{fig:cat-AZ-B}\n\\end{figure*}\n\n\nThe role of the symmetry energy on the properties of the crust at the neutron-drip transition can be understood as follows. Neglecting electron-ion interactions, \nthe neutron-drip condition~(\\ref{eq:n-drip-mue}) reduces to\n\\begin{equation}\n\\label{eq:mue-approx}\n\\mu_e\\approx \\mu_e^{\\rm drip} = m_e c^2 + \\frac{A}{Z}\\left(m_n c^2- \\frac{M^\\prime(A,Z)c^2}{A}\\right)\\, .\n\\end{equation}\nFor the sake of simplicity, let us consider a two-parameter mass formula: \n\\begin{equation}\n\\label{eq:2par-ldm}\nM^\\prime(A,Z)c^2=A \\left[a_{\\rm eff}+J_{\\rm eff} \\left(1-2 \\frac{Z}{A} \\right)^2+m_u c^2 \\right] + Z m_e c^2\\, ,\n\\end{equation}\nwhere $m_u$ is the unified mass unit, $a_{\\rm eff}<0$ is the contribution from charge-symmetric matter, while the deviations introduced by \nthe charge asymmetry are embedded in the coefficient $J_{\\rm eff}>0$. Note that due to nuclear surface effects the values of these coefficients \ndo not need to be the same as their corresponding values in infinite homogeneous nuclear matter at saturation. In particular, as already discussed \nin Sec.~\\ref{sec:hfb-models}, the ``effective'' symmetry energy coefficient $J_{\\rm eff}$ is expected to be smaller than $J$, and to decrease \nwith increasing $J$ or $L$. Minimizing $g$ using Eq.~(\\ref{eq:mue-approx}) and the mass formula~(\\ref{eq:2par-ldm}), the equilibrium proton fraction \nat the neutron-drip transition is approximately given by~\\cite{chamel2012} \n\\begin{equation}\n\\label{eq:zovera-cat}\n\\frac{Z}{A}\\approx \\frac{1}{2}\\sqrt{1+\\frac{a_{\\rm eff}}{J_{\\rm eff}}}\\, .\n\\end{equation}\nThis shows that $Z\/A$ decreases with increasing $L$ (decreasing $J_{\\rm eff}$). Note that in this analysis we have not made any assumption \nregarding the magnetic field. In other words, the correlation between $Z\/A$ and $L$ (or $J$) is thus expected to be independent of the magnetic \nfield strength (at least at the level of accuracy of the simple mass formula considered here), in agreement with the results plotted in the right panel of \nFig.~\\ref{fig:cat-AZ-B}. It follows from Eq.~(\\ref{eq:2par-ldm}) that decreasing $J_{\\rm eff}$ increases $M^\\prime(A,Z)$ (the energy cost associated \nwith charge asymmetry is reduced, therefore nuclei are more bound). Using Eq.~(\\ref{eq:mue-approx}), we find that $\\mu_e^{\\rm drip}$ increases with \n$L$. Since $n_{\\rm drip}$ and $P_{\\rm drip}$ increase with $\\mu_e^{\\rm drip}$, as shown in Eqs.~(\\ref{eq:ndrip-cat-approx}) and (\\ref{eq:Pdrip-cat-approx})\nin the absence of magnetic field, and in Eqs.~(\\ref{eq:ndrip-cat-approx-b}) and (\\ref{eq:Pdrip-cat-approx-b}) in the presence of a strongly quantizing \nmagnetic field, we can thus conclude that the neutron-drip transition is shifted to higher density and pressure with increasing the symmetry energy, as \nshown in Fig.~\\ref{fig:cat-ndrip-B}. \n\n\nThe equilibrium nucleus at the neutron-drip transition is less sensitive to the symmetry energy, as previously noticed in Ref.~\\cite{bao2014} in the absence \nof magnetic fields (see their Fig.~9). This can be understood as follows. The equilibrium with respect to weak interaction processes requires \n\\begin{equation}\n\\label{eq:betaeq}\n\\mu_p+\\mu_e=\\mu_n\\, ,\n\\end{equation}\nwhere $\\mu_p$ ($\\mu_n$) is the proton (neutron) chemical potential. Substituting the neutron-drip value of the neutron chemical potential $\\mu_n=m_n c^2$ in\nEq.~(\\ref{eq:betaeq}) and using Eq.~(\\ref{eq:mue-approx}), we obtain\n\\begin{equation}\n\\label{eq:mup}\n\\mu_p-m_p c^2 = Q_{n,\\beta}+\\frac{A}{Z}\\left(\\frac{M^\\prime(A,Z)c^2}{A}-m_n c^2 \\right)\\, ,\n\\end{equation}\nwhere $m_p$ is the proton mass and $Q_{n,\\beta} = 0.782$~MeV is the $\\beta$-decay energy of the neutron. The quantity on the left-hand side of Eq.~(\\ref{eq:mup}) \nis approximately equal to the opposite of the one-proton separation energy. This shows that the equilibrium nucleus is uniquely determined by nuclear masses only, \nand is sensitive to the details of the nuclear structure. As a consequence, the predicted nucleus depends on the nuclear mass model employed (see, e.g., \nRefs.~\\cite{roca2008,pearson2011,wolf2013,kreim2013,chamel2015c}). \nTo better illustrate this point, we have plotted in Fig.~\\ref{fig:mass_diff} the differences in \nthe mass predictions between HFB-22, HFB-25, and HFB-24 mass models for two isotopic chains, corresponding to the proton number at the neutron-drip \npoint (see also Table~\\ref{tab:drip-cat}). As shown in Fig.~\\ref{fig:mass_diff}, the HFB-22 model deviates more significantly from the ``reference'' \nmodel HFB-24 than HFB-25, thus explaining the quantitative differences in the dripping nucleus. The variations of $Z$ and $A$ with $L$ we find \nappear to be more irregular than those shown in Fig.~9 of Ref.~\\cite{bao2014}. This stems from the fact that in Ref.~\\cite{bao2014} nuclear masses were \ncalculated using the semi-classical Thomas-Fermi approximation, which does not take into account pairing and shell effects contrary to the fully quantum mechanical \nmass models~\\cite{goriely2013} employed here. \n\n\nThe presence of a strong magnetic field can change the composition at the neutron-drip point, as shown in the left panel of Fig.~\\ref{fig:cat-AZ-B} \n(see also Tables~\\ref{tab:drip-cat-mag-500}-\\ref{tab:drip-cat-mag-2000}). \nHowever, such behavior is only observed for the nuclear mass model HFB-22. In particular, the equilibrium nucleus is $^{122}$Kr for $B_\\star=0, 500$ and $1500$, \nwhile for $B_\\star=1000$ and $2000$ it is $^{128}$Sr. These results can be understood as follows. As discussed in Sec.~\\ref{sec:neutron-drip}, the equilibrium \nnucleus at the neutron-drip pressure $P_{\\rm drip}$ must be such as to minimize the Gibbs free energy per nucleon, therefore we must have \n\\begin{equation}\\label{eq:stability}\ng(A,Z,P_{\\rm drip}) < g(A^\\prime,Z^\\prime,P_{\\rm drip})\\, ,\n\\end{equation}\nfor any values of $A^\\prime\\neq A$ and $Z^\\prime\\neq Z$. This condition can be approximately expressed as~\\cite{chamel2015b}\n\\begin{equation}\\label{eq:dripping-nucleus}\n\\frac{M^\\prime(A^\\prime,Z^\\prime)c^2}{Z^\\prime} - \\frac{M^\\prime(A,Z)c^2}{Z} > \\left(\\frac{A^\\prime}{Z^\\prime}-\\frac{A}{Z}\\right) m_n c^2 + C e^2 n_e^{1\/3}\\left( Z^{2\/3}-Z^{\\prime\\, 2\/3}\\right)\\, ,\n\\end{equation}\nwhere the electron density $n_e$ has to be determined from Eq.~(\\ref{eq:n-drip-mue}). Equation~(\\ref{eq:dripping-nucleus}) can be equivalently written \nas $n_e < n_e^0$, where \n\\begin{equation}\\label{eq:ne0}\nn_e^0\\equiv \\biggl[\\frac{M^\\prime(A^\\prime,Z^\\prime)c^2}{Z^\\prime} - \\frac{M^\\prime(A,Z)c^2}{Z} - \\left(\\frac{A^\\prime}{Z^\\prime}-\\frac{A}{Z}\\right) m_n c^2\\biggr]^3\n\\biggl[C e^2 \\left( Z^{2\/3}-Z^{\\prime\\, 2\/3}\\right)\\biggr]^{-3}\\, .\n\\end{equation}\nThe HFB-22 nuclear mass model predicts very similar values for the \nthreshold electron Fermi energy $\\mu_e^{\\rm drip}$ for nuclei $^{128}$Sr and $^{122}$Kr: $24.970$ and $25.006$~MeV respectively. Substituting the theoretical \nvalues of the masses of $^{128}$Sr and $^{122}$Kr in Eq.~(\\ref{eq:ne0}) with $Z=36$, $A=122$, $Z^\\prime=38$, $A^\\prime=128$, we obtain $n_e^0\\approx 8.54\\times 10^{-5}$~fm$^{-3}$. \nDue to Landau quantization of electron motion, $n_e$ varies non-monotonically with $B_\\star$. As a consequence, the lattice term in Eq.~(\\ref{eq:dripping-nucleus}) \ncan thus become comparable to the other terms depending on $B_\\star$ to the effect that the condition (\\ref{eq:stability}) may be violated (i.e. $n_e\\geq n_e^0$), as \nshown in Fig.~\\ref{fig:drip-nuc-hfb22}. \nTransitions between $^{128}$Sr and $^{122}$Kr are found to occur at magnetic field strengths $B_\\star\\approx 861, 1239$ and $1883$. As shown in the right panel of \nFig.~\\ref{fig:cat-AZ-B}, the proton fraction $Z\/A$ is barely affected by these changes of composition. In other words, the correlation between $Z\/A$ and the symmetry \nenergy is almost independent of the magnetic field strength, as previously discussed. \n\n\n\\begin{figure*}\n\\begin{center}\n\\includegraphics[scale=0.45]{mass_differ_cat.eps}\n\\end{center}\n\\caption{(Color online) Difference in mass predictions for two pairs of Brussels-Montreal nuclear mass models along the two isotopic chains $Z=36$ and $Z=38$ relevant at neutron drip.}\n\\label{fig:mass_diff}\n\\end{figure*}\n\n\\begin{figure*}\n\\begin{center}\n\\includegraphics[scale=0.45]{drip_nuc_hfb22.eps}\n\\end{center}\n\\caption{Electron density $n_e$ (solid line) at the neutron-drip transition in nonaccreting neutron-star crusts as a \nfunction of the magnetic field strength, using the Brussels-Montreal nuclear mass model HFB-22 and assuming that the neutron-drip nucleus \nis $^{122}$Kr, as in the absence of magnetic field. The horizontal dotted line represents $n_e^0$, as given by Eq.~(\\ref{eq:ne0}). \nSee text for details. }\n\\label{fig:drip-nuc-hfb22}\n\\end{figure*}\n\n\n\\begin{table}\n\\centering\n\\caption{Neutron-drip transition in the crust of nonaccreting and unmagnetized neutron stars, as predicted by the HFB-22 to HFB-25 Brussels-Montreal nuclear mass models: mass and atomic numbers of the dripping nucleus, baryon density, and corresponding pressure. }\\smallskip\n\\label{tab:drip-cat}\n\\begin{tabular}{ccccc}\n\\hline\n & $A$ & $Z$ & $n_{\\rm drip}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip}$ ($10^{-4}$ MeV~fm$^{-3}$)\\\\\n\\hline \\noalign {\\smallskip}\nHFB-22 & 122 & 36 & 2.71 & 4.99 \\\\\nHFB-23 & 126 & 38 & 2.63 & 4.93 \\\\\nHFB-24 & 124 & 38 & 2.56 & 4.87 \\\\\nHFB-25 & 122 & 38 & 2.51 & 4.83 \\\\\n\\hline\n\\end{tabular}\n\\end{table}\n\n\n\n\\begin{table}\n\\centering\n\\caption{Neutron-drip transition in the crust of nonaccreting magnetized neutron stars with $B_\\star=500$, as predicted by the HFB-22 to HFB-25 Brussels-Montreal nuclear mass models: mass and atomic numbers of the dripping nucleus, baryon density, and corresponding pressure.}\\smallskip\n\\label{tab:drip-cat-mag-500}\n\\begin{tabular}{ccccc}\n\\hline\n & $A$ & $Z$ & $n_{\\rm drip}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip}$ ($10^{-4}$ MeV~fm$^{-3}$)\\\\\n\\hline \\noalign {\\smallskip}\nHFB-22 & 122 & 36 & 2.74 & 5.52 \\\\\nHFB-23 & 126 & 38 & 2.66 & 5.45 \\\\\nHFB-24 & 124 & 38 & 2.61 & 5.39 \\\\\nHFB-25 & 122 & 38 & 2.55 & 5.35 \\\\\n\\hline\n\\end{tabular}\n\\end{table}\n\n\\begin{table}\n\\centering\n\\caption{Same as in Table~\\ref{tab:drip-cat-mag-500} but for $B_\\star=1000$.}\\smallskip\n\\label{tab:drip-cat-mag-1000}\n\\begin{tabular}{ccccc}\n\\hline\n & $A$ & $Z$ & $n_{\\rm drip}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip}$ ($10^{-4}$ MeV~fm$^{-3}$)\\\\\n\\hline \\noalign {\\smallskip}\nHFB-22 & 128 & 38 & 3.06 & 6.70 \\\\\nHFB-23 & 126 & 38 & 2.98 & 6.63 \\\\\nHFB-24 & 124 & 38 & 2.91 & 6.56 \\\\\nHFB-25 & 122 & 38 & 2.85 & 6.52 \\\\\n\\hline\n\\end{tabular}\n\\end{table}\n\n\\begin{table}\n\\centering\n\\caption{Same as in Table~\\ref{tab:drip-cat-mag-500} but for $B_\\star=1500$.}\\smallskip\n\\label{tab:drip-cat-mag-1500}\n\\begin{tabular}{ccccc}\n\\hline\n & $A$ & $Z$ & $n_{\\rm drip}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip}$ ($10^{-4}$ MeV~fm$^{-3}$)\\\\\n\\hline \\noalign {\\smallskip}\nHFB-22 & 122 & 36 & 2.30 & 8.66 \\\\\nHFB-23 & 126 & 38 & 2.24 & 8.60 \\\\\nHFB-24 & 124 & 38 & 2.20 & 8.56 \\\\\nHFB-25 & 122 & 38 & 2.16 & 8.52 \\\\\n\\hline\n\\end{tabular}\n\\end{table}\n\n\\begin{table}\n\\centering\n\\caption{Same as in Table~\\ref{tab:drip-cat-mag-500} but for $B_\\star=2000$.}\\smallskip\n\\label{tab:drip-cat-mag-2000}\n\\begin{tabular}{ccccc}\n\\hline\n & $A$ & $Z$ & $n_{\\rm drip}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip}$ ($10^{-3}$ MeV~fm$^{-3}$)\\\\\n\\hline \\noalign {\\smallskip}\nHFB-22 & 128 & 38 & 3.06 & 11.6 \\\\\nHFB-23 & 126 & 38 & 3.00 & 11.6 \\\\\nHFB-24 & 124 & 38 & 2.95 & 11.5 \\\\\nHFB-25 & 122 & 38 & 2.89 & 11.4 \\\\\n\\hline\n\\end{tabular}\n\\end{table}\n\n\n\\subsection{Accreting neutron stars}\n\\label{sec:results-accreting}\n\nFor accreting neutron-star crusts, we have studied the neutron-drip transition as explained in Sec.~\\ref{sec:neutron-drip} (see also Ref.~\\cite{chamel2015a}). \nWe have considered different initial compositions: the ashes produced by an $rp$-process during an x-ray burst \\cite{schatz2001}, and the ashes produced by \nsteady state hydrogen and helium burning \\cite{schatz2003} as expected to occur during superbursts \\cite{gupta2007}. After determining the dripping nucleus, we \nhave calculated the neutron-drip density and pressure by solving numerically Eq.~(\\ref{eq:e-capture+n-emission-gibbs-approx}) considering all possible \nneutron-emission processes. Results are summarized in Tables~\\ref{tab:drip-acc-22}-\\ref{tab:drip-acc-25} for different nuclear mass models. \nFor comparison with previous works~\\cite{hz1990, hz2003}, we have also considered ashes of x-ray bursts consisting of pure $^{56}$Fe. Results are indicated in \nTable~\\ref{tab:drip-acc-a56}. \n\nAs already pointed out in Ref.~\\cite{chamel2015a}, for a given nuclear mass model the neutron-drip transition in accreting neutron stars can occur at \neither lower or at higher densities and pressures than in nonaccreting neutron stars. Depending on the mass model adopted, the neutron-drip density thus \nranges from $1.60 \\times 10^{-4}$~fm$^{-3}$ to $3.90 \\times 10^{-4}$~fm$^{-3}$, \nand the corresponding pressure from $2.77 \\times 10^{-4}$~MeV~fm$^{-3}$ to $7.77 \\times 10^{-4}$~MeV~fm$^{-3}$. The numerical results \nobtained solving Eq.~(\\ref{eq:e-capture+n-emission-gibbs-approx}) are reproduced by the analytical formulas~(\\ref{eq:ndrip-acc}) and \n(\\ref{eq:Pdrip-acc}) with an error of $0.1\\%$ at most. \n\nThe change of the neutron-drip density with the slope $L$ of the symmetry energy is found to be very different from that obtained in \nnonaccreting neutron-star crusts. As shown in Fig.~\\ref{fig:acc-ndrip}, for some ashes like $^{104}$Cd, no obvious correlation is observed while for \nother ashes like $^{66}$Ni, $n_{\\rm drip-acc}$ appears to be \\emph{anticorrelated} with $L$: $n_{\\rm drip-acc}$ decreases with increasing $L$. \nThis behavior can be understood as follows. Ignoring electron-ion interactions, the neutron-drip condition~(\\ref{eq:e-capture+n-emission-gibbs-approx}) \nreduces to\n\\begin{equation}\n\\mu_e\\approx \\mu_e^{\\rm drip-acc} = M^\\prime(A-\\Delta N,Z-1)c^2-M^\\prime(A,Z)c^2 +m_n c^2 \\Delta N + m_e c^2\\, , \n\\end{equation}\nwhich can be more conveniently written as \n\\begin{equation}\\label{eq:muedrip-acc-s}\n\\mu_e^{\\rm drip-acc} = S_{\\Delta N n}(A,Z-1) + \\mu_e^\\beta(A,Z)\\, ,\n\\end{equation}\nusing Eq.~(\\ref{eq:sn-dn}) and introducing the threshold electron Fermi energy for the onset of electron captures (see, e.g. Ref.~\\cite{chamel2015d} for a recent \ndiscussion)\n\\begin{equation}\n\\mu_e^\\beta(A,Z)=M^\\prime(A,Z-1)c^2 -M^\\prime(A,Z)c^2 + m_e c^2\\, .\n\\end{equation}\nThe mass difference $\\Delta M^\\prime=M'(A,Z-1) -M'(A,Z)$, which represents the change of mass associated with the substitution of a proton by a neutron, \nis expected to be mainly determined by symmetry energy effects. On the other hand, the $\\Delta N$-neutron separation energy $S_{\\Delta N n}(A,Z-1)$ \nis likely to be more dependent on the details of the nuclear structure than on the symmetry energy. As shown in Tables~\\ref{tab:drip-acc-22}-\\ref{tab:drip-acc-a56}, \n$|S_{\\Delta N n}(A,Z-1)| \\ll \\Delta M^\\prime c^2$ therefore $\\mu_e^{\\rm drip-acc}\\approx \\mu_e^\\beta$. On the other hand, as discussed in Sec.~\\ref{sec:neutron-drip}, \nthe composition of accreting neutron-star crusts at the neutron-drip transition is directly determined by the condition $S_{\\Delta N n}(A,Z-1)<0$. Provided \nthe dependence on the symmetry energy of $S_{\\Delta N n}(A,Z-1)$ is weak enough, the dripping nucleus will thus be independent of $L$. In this case, the \nvariations of $S_{\\Delta N n}(A,Z-1)$ with $L$ are typically much smaller than the variations of $\\Delta M^\\prime c^2$, as shown in Figs.~\\ref{fig:deltamprime} and \n\\ref{fig:Sdeltan}. Using the simple mass formula~(\\ref{eq:2par-ldm}), we find\n\\begin{equation}\n\\mu_e^\\beta(A,Z)=\\Delta M^\\prime c^2 +m_e c^2 \\approx 4J_{\\rm eff}\\left(1+\\frac{1-2Z}{A}\\right)\\, .\n\\end{equation}\nThis means that with increasing $J$ or $L$ (decreasing $J_{\\rm eff}$), $\\Delta M^\\prime$ and $\\mu_e^\\beta(A,Z)$ both \\emph{decrease}. It thus follows from \nEqs.~(\\ref{eq:ndrip-acc}), (\\ref{eq:Pdrip-acc}), and Eqs.~(\\ref{eq:muedrip-acc-s}), that $\\mu_e^{\\rm drip-acc}$, $n_{\\rm drip-acc}$ and $P_{\\rm drip-acc}$ \nalso decrease with $L$, as shown in the upper panel of Fig.~\\ref{fig:acc-ndrip}. \nIn the peculiar case of $A=105$, the rather low value of $\\Delta M^\\prime c^2$ predicted by HFB-25 is compensated by a comparatively high value $S_{\\Delta N n}(A,Z-1)$, \nas can be seen in Figs.~\\ref{fig:deltamprime} and \\ref{fig:Sdeltan}. As a result, $n_{\\rm drip-acc}$ is still found to decrease with increasing $L$ despite the nonmonotonic \nvariation of $\\Delta M^\\prime$, as shown in Fig.~\\ref{fig:acc-ndrip}. \nFor some ashes, the variations of $S_{\\Delta N n}(A,Z-1)$ are comparable to those of $\\Delta M^\\prime c^2$, and large enough to even change the composition. This \nleads to nonmonotonic variations of the neutron-drip density and pressure, as illustrated in the lower panel of Fig.~\\ref{fig:acc-ndrip}. In these cases, \neffects other than the symmetry energy play a role. The anticorrelation between $n_{\\rm drip-acc}$ (or $P_{\\rm drip-acc}$) and $L$ thus relies to \na large extent on the importance of nuclear structure effects far from the stability valley. \n\n\\begin{table}\n\\centering\n\\caption{Neutron-drip transition in the crust of accreting neutron stars, as predicted by the HFB-22 Brussels-Montreal nuclear mass model: mass and atomic numbers of the dripping nucleus, number of emitted neutrons, baryon density $n_{\\rm drip-acc}$ ($10^{-4}$ fm$^{-3}$), and corresponding pressure $P_{\\rm drip-acc}$ ($10^{-4}$ MeV~fm$^{-3}$), $S_{\\Delta N n}(A,Z-1)$ (MeV), $\\Delta M^\\prime$ (MeV$\/c^2$), and $\\mu_e^{\\rm drip-acc}$ (MeV).\nThe mass numbers $A$ are listed from top to bottom considering that the ashes are produced by ordinary\nx-ray bursts (upper panel) or superbursts (lower panel). See text for details.}\\smallskip\n\\label{tab:drip-acc-22}\n\\begin{tabular}{cccccccc}\n\\hline\n $A$ & $ Z$ & $\\Delta N$& $n_{\\rm drip-acc}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip-acc}$ ($10^{-4}$ MeV~fm$^{-3}$) & $S_{\\Delta N n}$ (MeV) & $\\Delta M^\\prime$ (MeV$\/c^2$) & $\\mu_e^{\\rm drip-acc}$ (MeV) \\\\\n \\hline\n 104 & 32 & 1 & 2.71 & 5.31 & -0.79 & 25.14 & 24.86 \\\\\n 105 & 33 & 1 & 1.90 & 3.40 & -1.00 & 22.70 & 22.21 \\\\\n 68 & 22 & 1 & 2.31 & 4.64 & -0.28 & 24.14 & 24.37 \\\\\n 64 & 18 & 5 & 3.90 & 7.77 & -1.85 & 29.23 & 27.89 \\\\\n 72 & 22 & 1 & 2.89 & 5.78 & -0.31 & 25.55 & 25.75 \\\\\n 76 & 24 & 1 & 2.86 & 5.95 & -0.17 & 25.52 & 25.86 \\\\\n 98 & 32 & 1 & 1.93 & 3.66 & -0.20 & 22.34 & 22.65 \\\\\n 103 & 33 & 1 & 1.60 & 2.77 & -0.02 & 20.62 & 21.11 \\\\\n 106 & 32 & 1 & 2.92 & 5.71 & -0.69 & 25.50 & 25.32 \\\\\n \\hline\n 66 & 22 & 1 & 1.99 & 3.95 & -0.19 & 23.09 & 23.41 \\\\\n 64 & 18 & 5 & 3.90 & 7.77 & -1.85 & 29.23 & 27.89 \\\\\n 60 & 20 & 1 & 1.83 & 3.55 & -1.67 & 24.02 & 22.86 \\\\\n \\end{tabular}\n \\end{table}\n\n\\begin{table}\n\\centering\n\\caption{Same as in Table~\\ref{tab:drip-acc-22} but for the HFB-23 Brussels-Montreal nuclear mass model.}\\smallskip\n\\label{tab:drip-acc-23}\n\\begin{tabular}{cccccccc}\n\\hline\n $A$ & $Z$ & $\\Delta N$& $n_{\\rm drip-acc}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip-acc}$ ($10^{-4}$ MeV~fm$^{-3}$) & $S_{\\Delta N n}$ (MeV) & $\\Delta M^\\prime$ (MeV$\/c^2$) & $\\mu_e^{\\rm drip-acc}$ (MeV) \\\\\n \\hline\n 104 & 32 & 1 & 2.83 & 5.62 & -1.02 & 25.73 & 25.22 \\\\\n 105 & 33 & 1 & 1.97 & 3.57 & -1.33 & 23.30 & 22.48 \\\\\n 68 & 22 & 1 & 2.35 & 4.73 & -0.56 & 24.54 & 24.49 \\\\\n 64 & 20 & 1 & 3.27 & 7.04 & -0.13 & 26.75 & 27.13 \\\\\n 72 & 22 & 1 & 3.06 & 6.24 & -0.11 & 25.84 & 26.24\\\\\n 76 & 24 & 1 & 2.94 & 6.18 & -0.38 & 25.98 & 26.11 \\\\\n 98 & 32 & 1 & 1.98 & 3.77 & -0.50 & 22.82 & 22.83 \\\\\n 103 & 31 & 1 & 2.45 & 4.50 & -0.91 & 24.29 & 23.89 \\\\\n 106 & 34 & 1 & 2.11 & 4.02 & -0.009 & 22.63 & 23.13 \\\\\n \\hline\n 66 & 22 & 1 & 2.07 & 4.16 & -0.23 & 23.43 & 23.71 \\\\\n 64 & 20 & 1 & 3.27 & 7.04 & -0.13 & 26.75 & 27.13 \\\\\n 60 & 20 & 1 & 1.93 & 3.79 & -1.77 & 24.50 & 23.24 \\\\\n \\end{tabular}\n \\end{table}\n\n\\begin{table}\n\\centering\n\\caption{Same as in Table~\\ref{tab:drip-acc-22} but for the HFB-24 Brussels-Montreal nuclear mass model.}\\smallskip\n\\label{tab:drip-acc-24}\n\\begin{tabular}{cccccccc}\n\\hline\n $A$ & $Z$ & $\\Delta N$& $n_{\\rm drip-acc}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip-acc}$ ($10^{-4}$ MeV~fm$^{-3}$) & $S_{\\Delta N n}$ (MeV) & $\\Delta M^\\prime$ (MeV$\/c^2$) & $\\mu_e^{\\rm drip-acc}$ (MeV) \\\\\n \\hline\n 104 & 32 & 1 & 2.87 & 5.73 & -1.49 & 26.32 & 25.34 \\\\\n 105 & 33 & 1 & 2.10 & 3.89 & -1.11 & 23.57 & 22.97 \\\\\n 68 & 22 & 1 & 2.45 & 5.00 & -0.75 & 25.07 & 24.83 \\\\\n 64 & 22 & 1 & 1.66 & 3.22 & -0.07 & 21.81 & 22.25 \\\\\n 72 & 22 & 1 & 3.08 & 6.28 & -0.29 & 26.07 & 26.29 \\\\\n 76 & 24 & 1 & 3.10 & 6.61 & -0.46 & 26.50 & 26.55 \\\\\n 98 & 32 & 1 & 2.04 & 3.94 & -0.38 & 22.95 & 23.08 \\\\\n 103 & 31 & 3 & 2.49 & 4.59 & -0.89 & 24.39 & 24.01 \\\\\n 106 & 34 & 1 & 2.17 & 4.16 & -1.10 & 23.92 & 23.33 \\\\\n \\hline\n 66 & 22 & 1 & 2.09 & 4.22 & -0.29 & 23.58 & 23.80 \\\\\n 64 & 22 & 1 & 1.66 & 3.22 & -0.07 & 21.81 & 22.25 \\\\\n 60 & 20 & 1 & 2.03 & 4.05 & -1.66 & 24.78 & 23.63 \\\\\n \\end{tabular}\n \\end{table}\n\n\\begin{table}\n\\centering\n\\caption{Same as in Table~\\ref{tab:drip-acc-22} but for the HFB-25 Brussels-Montreal nuclear mass model.}\\smallskip\n\\label{tab:drip-acc-25}\n\\begin{tabular}{cccccccc}\n\\hline\n $A$ & $Z$ & $\\Delta N$& $n_{\\rm drip-acc}$ ($10^{-4}$ fm$^{-3}$) & $P_{\\rm drip-acc}$ ($10^{-4}$ MeV~fm$^{-3}$) & $S_{\\Delta N n}$ (MeV) & $\\Delta M^\\prime$ (MeV$\/c^2$) & $\\mu_e^{\\rm drip-acc}$ (MeV) \\\\\n \\hline\n 104 & 34 & 1 & 2.02 & 3.87 & -0.009 & 22.42 & 22.92 \\\\\n 105 & 33 & 1 & 2.18 & 4.07 & -0.57 & 23.30 & 23.24 \\\\\n 68 & 22 & 1 & 2.59 & 5.40 & -0.75 & 25.55 & 25.31 \\\\\n 64 & 22 & 1 & 1.79 & 3.59 & -0.18 & 22.52 & 22.85 \\\\\n 72 & 22 & 1 & 3.27 & 6.82 & -0.55 & 26.87 & 26.83 \\\\\n 76 & 24 & 1 & 3.14 & 6.73 & -0.58 & 26.74 & 26.67 \\\\\n 98 & 32 & 1 & 2.11 & 4.12 & -0.64 & 23.47 & 23.34 \\\\\n 103 & 33 & 1 & 1.78 & 3.18 & -0.76 & 22.10 & 21.85 \\\\\n 106 & 34 & 1 & 2.25 & 4.36 & -0.05 & 23.15 & 23.61 \\\\\n \\hline\n 66 & 22 & 1 & 2.20 & 4.52 & -0.22 & 23.92 & 24.21 \\\\\n 64 & 22 & 1 & 1.79 & 3.59 & -0.18 & 22.52 & 22.85 \\\\\n 60 & 20 & 1 & 2.08 & 4.21 & -1.94 & 25.28 & 23.85 \\\\\n \\end{tabular}\n \\end{table}\n\n\n\\begin{table}\n\\centering\n\\caption{Neutron-drip transition in the crust of accreting neutron stars, as predicted by different Brussels-Montreal nuclear mass models for $^{56}$Fe ashes: atomic number $Z$ of the dripping nucleus, number of emitted neutrons, density and corresponding pressure, $S_{\\Delta N n}(A,Z-1)$, $\\Delta M^\\prime$ (MeV$\/c^2$), and $\\mu_e^{\\rm drip-acc}$ (MeV). See text for details.}\\smallskip\n\\label{tab:drip-acc-a56}\n\\begin{tabular}{ccccc}\n\\hline\n & HFB-22 & HFB-23 & HFB-24 & HFB-25 \\\\\n\\hline\n$Z$ & 18 & 18 & 18 & 18 \\\\\n$\\Delta N$ & 1 & 1 & 3 & 1 \\\\\n$n_{\\rm drip-acc}$ ($10^{-4}$ fm$^{-3}$) & 2.49 & 2.58 & 2.73 & 2.84 \\\\\n$P_{\\rm drip-acc}$ ($10^{-4}$ MeV~fm$^{-3}$) & 5.10 & 5.34 & 5.77 & 6.07 \\\\\n$S_{\\Delta N n}$ (MeV) & -1.17 & -1.32 & -3.27 & -1.61 \\\\\n$\\Delta M^\\prime$ (MeV$\/c^2$) & 25.76 & 26.20 & 28.64 & 27.32 \\\\\n$\\mu_e^{\\rm drip-acc}$ (MeV) & 25.10 & 25.39 & 25.88 & 26.22 \\\\\n\\end{tabular}\n\\end{table}\n\n\n\n\\begin{figure*}\n\\begin{center}\n\\includegraphics[scale=0.45]{drip_acc_22-25_ndrip_new.eps}\n\\end{center}\n\\caption{(Color online) Neutron-drip density as a function of the slope $L$ of the symmetry energy of infinite homogeneous nuclear matter at saturation, as obtained using the HFB-22 to HFB-25 Brussels-Montreal nuclear mass \nmodels, for accreting neutron-star crusts with different initial composition of ashes (see text for details).}\n\\label{fig:acc-ndrip}\n\\end{figure*}\n\n\\begin{figure*}\n\\begin{center}\n\\includegraphics[scale=0.45]{drip_acc_22-25_deltam.eps}\n\\end{center}\n\\caption{(Color online) Mass difference $\\Delta M^\\prime$ (in units of MeV$\/c^2$) as a function of the slope $L$ of the symmetry energy of infinite homogeneous nuclear matter at saturation, as obtained using the \nHFB-22 to HFB-25 Brussels-Montreal nuclear mass models, for accreting \nneutron-star crusts with different initial composition of ashes (see text for details).}\n\\label{fig:deltamprime}\n\\end{figure*}\n\n\\begin{figure*}\n\\begin{center}\n\\includegraphics[scale=0.45]{drip_acc_22-25_Sdeltan.eps}\n\\end{center}\n\\caption{(Color online) $\\Delta N$-neutron separation energy $S_{\\Delta N n}$ (in units of MeV) as a function of the slope $L$ of the symmetry energy of infinite homogeneous nuclear matter at saturation, as obtained using the \nHFB-22 to HFB-25 Brussels-Montreal nuclear mass models, for accreting \nneutron-star crusts with different initial composition of ashes (see text for details).}\n\\label{fig:Sdeltan}\n\\end{figure*}\n\n\n\\section{Conclusions}\n\\label{sec:conclusions}\n\nWe have studied the role of the symmetry energy on the neutron-drip transition in both accreting and nonaccreting neutron-star crusts. We have also allowed for \nthe presence of a strong magnetic field, as in magnetars. The masses of nuclei encountered in this region of the neutron-star crust are experimentally unknown. \nFor this reason, we have employed a recent family of microscopic nuclear mass models, from HFB-22 to HFB-25, developed by the Brussels-Montreal collaboration~\\cite{goriely2013}. These models provide equally good fits to the 2353 measured masses of nuclei with $N$ and $Z \\geq 8$ from the 2012 Atomic Mass \nEvaluation~\\cite{audi2012}, with a root-mean-square deviation of about $0.6$~MeV. \nOn the other hand, these models lead to different predictions for the behavior of the symmetry energy in infinite homogeneous\nnuclear matter.\nIn particular, these functionals \nwere constrained so as to yield different \nvalues of the symmetry energy at saturation, from $J=29$~MeV to $J=32$~MeV, the slope of the symmetry energy ranging from $L=37$~MeV to $L=69$~MeV. \n\nFor nonaccreting weakly magnetized neutron stars, the neutron-drip density $n_{\\rm drip}$ is found to increase almost linearly with $L$ (or equivalently with $J$) \nwhile the proton fraction $Z\/A$ decreases, in agreement with previous studies~\\cite{bao2014} (see also Refs.~\\cite{roca2008,provid2014,grill2014}). \nIn the presence of a strong magnetic field, the dripping nucleus\nhence also $Z\/A$ is unchanged, as found in Refs.~\\cite{chamel2012,chamel2015b}, for all models but HFB-22. In this case, the dripping nucleus alternates between \n$^{122}$Kr and $^{128}$Sr depending on the magnetic field strength. This peculiar behavior arises from Landau quantization of electron motion and from the fact \nthat the threshold electron Fermi energy $\\mu_e^{\\rm drip}$ are almost equal. Their proton fraction are also very similar so that all in all the linear \ncorrelation between $Z\/A$ and $L$ is hardly affected by the magnetic field. The neutron-drip density $n_{\\rm drip}$ exhibits typical quantum oscillations as \na function of the magnetic field strength, as recently discussed in Ref.~\\cite{chamel2015b}. Still, $n_{\\rm drip}$ remains linearly correlated with $L$. \nAlthough a soft symmetry energy favors neutron drip in isolated nuclei~\\cite{todd2003}, this result does not necessarily imply the observed correlation \nbetween $n_{\\rm drip}$ and $L$. Indeed, as recently discussed in Ref.~\\cite{chamel2015a}, the dripping nucleus in the crust is actually stable \nagainst neutron emission, but unstable against electron captures followed by neutron emission. In fact, such a correlation is not found in accreting neutron-star crusts. \nDepending on the initial composition of the ashes from x-ray bursts and superbursts, $n_{\\rm drip}$ \n\\emph{decreases} almost linearly with increasing $L$ while the dripping nucleus remains the same. In other cases, the symmetry energy does not seem \nto play any role. \n\nWe have qualitatively explained these different behaviors using a simple mass formula, and making use of the analytical expressions for the neutron-drip density and \npressure obtained in Refs.~\\cite{chamel2015a,chamel2015b}. In particular, we have shown that the anticorrelation between $n_{\\rm drip}$ and $L$ in accreting \nneutron stars depends to a large extent to the relative importance of nuclear structure effects (like shell effects and pairing) and symmetry energy effects on the \nneutron separation energy. More precisely, the anticorrelation is broken whenever the differences between the neutron separation energies predicted by the different \nmass models are large enough to change the dripping nucleus. \n\nIn any case, the composition of the deepest layers of the outer crust of a neutron star is very sensitive to the details of the nuclear structure far from the stability valley. In nonaccreting neutron-star crusts, the neutron-drip transition is mainly governed by the values of the masses of very neutron-rich strontium and krypton isotopes. \nAlthough the masses of these nuclei have not yet been measured, the composition of nonaccreting neutron-star crusts has been recently constrained by experiment \nto deeper layers~\\cite{wolf2013}. The nuclei thought to be present in accreting neutron star crusts span a much larger region of the nuclear chart, depending on \nthe ashes from x-ray bursts and superbursts. In these neutron stars, the neutron-drip transition is not directly determined by nuclear masses but rather by some \ncombinations of masses, namely the (multiple) neutron separation energies and the isobaric two-point mass differences. \n\nThe onset of neutron emission by nuclei marks the transition to the inner region of the neutron-star crust, where neutron-proton clusters coexist with a neutron liquid. \nIn turn, this neutron liquid, which becomes superfluid at low enough temperatures, is expected to play a role in various observed astrophysical phenomena (see, e.g., \nRef.~\\cite{chamelhaensel2008} for a review) like sudden spin-ups and spin-downs (so-called ``glitches'' and ``antiglitches'', respectively)~\\cite{dib2008, gug2014, \narchibald2013, sasmaz2014, duncan2013, kantor2014}, quasiperiodic oscillations detected in the giant flares from soft $\\gamma$-ray repeaters~\\cite{passamonti2014}, \ncooling of strongly magnetized neutron stars~\\cite{aguilera2009}, deep crustal heating (most of the heat being released near the neutron-drip transition~\\cite{haensel2008}), \nand the thermal relaxation of quasipersistent soft x-ray transients~\\cite{shternin2007,brown2009,page2013}. By shifting the neutron-drip transition to higher or lower densities, the symmetry energy may thus leave its imprint on these astrophysical phenomena. \n\n\n\n\\begin{acknowledgments}\nThis work has been mainly supported by Fonds de la Recherche Scientifique - FNRS (Belgium), and by the bilateral project between Fonds de la Recherche Scientifique - FNRS (Belgium), Wallonie-Bruxelles-International (Belgium) and the Bulgarian Academy of Sciences. \nThis work has been also partially supported by the Bulgarian National Science Fund under Contract No. DFNI-T02\/19t and the Cooperation in Science and Technology (COST Action) MP1304 ``NewCompStar''. The authors would like to thank J.~M. Pearson for very fruitful discussions.\n\\end{acknowledgments}\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}