{"text":"\\section{Introduction}\n\\label{sec:introduction}\nQuivers with relations play a fundamental role in the representation theory of finite dimensional algebras.\n\nEvery basic finite dimensional algebra $\\Lambda$ over an algebraically closed field $\\Bbbk$ is isomorphic to a path algebra $\\Bbbk Q\/R$ of a finite quiver $Q$ with admissible relations $R$. Moreover, the modules of $\\Lambda$ are in bijection with the representations of $(Q,I)$ over $\\Bbbk$. Without relations, we get a correspondence only to hereditary algebras \\cite{Gabriel}; see also \\cite[Ch.~III.1]{ARS}. \n\nNon-hereditary algebras are central in most fields of modern representation theory of algebras. For one, higher homological algebra requires algebras of global dimension at least 2 \\cite{Jasso, Kvamme}. There is a rich tradition of studying classes of non-hereditary algebras, such as gentle \\cite{AH, AS}, clannish \\cite{C-B}, Schur \\cite{Erdmann}, preprojective \\cite{Ringel}, and self--injective \\cite{SY} algebras.\n\nContinuous quivers and their representations were first explicitly studied in \\cite{IRT}. They are a natural generalisation of quivers, replacing finite sets of vertices with uncountably infinite sets. In the process, one gains intuition about what characteristics of representation theory come from innate properties of algebraic structures, and what comes from the discrete examples that are usually studied.\n\nOne parameter persistence modules are often defined over the real line so that persistence modules coincide with pointwise finite-dimensional representations of a continuous quiver of type $\\mathbb{A}$ (see, for example, \\cite{CdSGO}).\nIn \\cite{BBOS} the authors consider $m\\times n$ rectangular grid quivers which have the commutativity relation on each square.\nThe authors of \\cite{BBH} study homological approximations in order to obtain new invariants of these representations (persistence modules).\n\n\nGiven the important role of quiver relations in the representation theory of finite-dimensional algebra, it is natural to ask if relations can be extended to the continuous setting. This has already been done in a restricted sense by the second author and Zhu in \\cite{RZ}. We give a more general definition that works with any underlying quiver. To capture the full generality we actually go beyond quivers and consider categories instead.\n\nIn starting this work, we were motivated by two areas of study that we intend to lift to the continuous setting: gentle algebras and $d$-cluster-tilting subcategories. In gentle algebras, the relations appear in the definition, and are always generated by compositions of two arrows. This type of relations are generalized as \\emph{point relations} in \\Cref{subsec:point relations}. An important class of $d$-cluster-tilting subcategories appear in the module category of type A algebras, with relations consisting of all paths above a certain length \\cite{Vaso}. This type of relations is generalized as \\emph{length relations} in \\Cref{subsec:length}.\n\n\n\\subsection{Contributions}\nIn \\Cref{sec:definition and general}, we give essential background, before stating our main definition.\n\n\\begin{definition}[\\Cref{def:admissible}]\nLet $\\mathcal{C}$ be a category and $\\mathcal{I}$ an ideal in $\\mathcal{C}$.\n We say $\\mathcal{I}$ is \\emph{admissible} if the following are satisfied.\n \\begin{enumerate}\n \\item For each $f$ in $\\mathcal{I}$, there exists a finite collection of morphisms $g_1,\\ldots,g_n$ not in $\\mathcal{I}$ such that $f = g_n\\circ\\cdots\\circ g_1$.\n \\item For each nonzero endomorphism $f$, if $f$ is not an isomorphism then there exists $n\\geq 2\\in\\mathbb{N}$ such that $f^n\\in\\mathcal{I}(X,X)$.\n \\end{enumerate}\n\\end{definition}\n\n\\Cref{sec:general results} gives some general results on the quotient category $\\mathcal{C} \/ \\mathcal{I}$, summarized here.\n\n\n\\begin{theorem}[\\Cref{prop:connected,prop:basic,prop:radical commutes with admissible}]\n Let $\\mathcal{C}$ be a category. Let $\\mathcal{I}$ be an admissible ideal of morphisms.\n \n \\begin{enumerate}\n \\item If $\\mathcal{C}$ is connected, then $\\mathcal{C} \/ \\mathcal{I}$ is also connected.\n \\item If $\\mathcal{C}$ is Krull--Remak--Schmidt--Azumaya, then $\\mathcal{C} \/ \\mathcal I$ is also Krull--Remak--Schmidt--Azumaya.\n \\item If all endomorphism rings of $\\mathcal{C}$ are artinian, then \\[\\mathsf{Rad}(\\mathcal{C} \/ \\mathcal{I})=\\mathsf{Rad}(\\mathcal{C}) \/ \\mathcal{I}.\\]\n \\end{enumerate}\n\\end{theorem}\n\nIn \\Cref{sec:relations} we give two important classes of relations that can generate admissible ideals. The first is \\emph{point relations}, which generalizes relations of length two. The idea is that certain paths through a vertex in the quiver are excluded, but others not. For an illustration see \\Cref{fig:point intro}.\n\n\\begin{figure}[h!]\n \\centering\n \\begin{subfigure}[b]{0.3\\textwidth}\n \\centering\n \\begin{tikzpicture}\n \\node (center) at (0,0) {$\\bullet$};\n \\node at (0,.3) {$P$};\n \\node (left) at (-1,0) {$\\circ$};\n \\node (right) at (1,0) {$\\circ$};\n \\draw[->] (left)--(center);\n \\draw[->] (center) -- (right);\n \\end{tikzpicture}\n \\caption{Discrete case}\n \\end{subfigure}\n \\begin{subfigure}[b]{0.3\\textwidth}\n \\centering\n \\begin{tikzpicture}[decoration={markings, mark=at position 0.5 with {\\arrow{>}}}]\n \\node (center) at (0,0) {$\\bullet$};\n \\node (left) at (-1,0) {};\n \\node (right) at (1,0) {};\n \\node at (0,.3) {$P$};\n \\draw[very thick, postaction=decorate] (left.center) to[bend left] (center.center);\n \\draw[very thick, postaction=decorate] (center.center) to[bend right] (right.center);\n \\end{tikzpicture}\n \\caption{Continuous case}\n \\end{subfigure}\n \\caption{An illustration of point relations in the discrete and continuous case. In both cases, the relations contain all paths passing through the point $P$.\n See \\Cref{fig:arrow and line} on page~\\pageref{fig:arrow and line} for an explanation of our drawing conventions.}\n \\label{fig:point intro}\n\\end{figure}\n\n\n\\begin{theorem}[\\Cref{thm:point relations are admissible}]\n Let $\\{\\mathcal P_\\alpha\\}$ be an admissible collection of point relations in $\\mathcal{C}$, such that any cycles are either isomorphisms (and hence trivial) or contained in at least one $\\mathcal P_\\alpha$.\n Then $\\mathcal{I}=\\langle \\bigcup_\\alpha \\mathcal P_\\alpha \\rangle$ is an admissible ideal in $\\mathcal{C}$.\n\\end{theorem}\n\nThe other class of relations we define are \\emph{length relations}. This is a generalisation of relations generated by paths containing at least $n$ arrows, where $n$ is a natural number.\n\n\\begin{theorem}[\\Cref{thm:length relations are admissible}]\n A length relation generates an admissible ideal.\n\\end{theorem}\n\\Cref{sec:examples} contains multiple examples of how relations work, including a sketch of their Auslander--Reiten theory.\n\n\\subsection{Future Work}\nThe present paper is a precursor to future work on generalizations of non-hereditary structures.\nOf note, the authors will consider point relations, such as \\Cref{ex:crossing real lines}, that generalize gentle algebras. They will also study the modding out by length relations, such as \\Cref{ex:length relations}(\\ref{ex:length relations:continuous A}), to generate higher cluster tilting subcategories.\n\n\\subsection{Acknowledgements}\nThe idea for this project was conceived at the Hausdorff Research Institute of Mathematics, KMJ visited JDR at Ghent University during this project, and JDR visited KMJ at Aarhus University during this project.\nThe authors thank each of these institutions for their hospitality.\nKMJ is supported by the Norwegian Research Council via the project Higher Homological Algebra and Tilting Theory (301046).\nJDR is supported at Ghent University by BOF grant 01P12621.\nThe authors would like to thank Jenny August, Raphael Bennett-Tennenhaus, Charles Paquette, Amit Shah, Emine Y{\\i}ld{\\i}r{\\i}m, and Shijie Zhu for helpful discussions.\n\n\\subsection{Conventions}\nWe work over $\\Bbbk=\\overline{\\Bbbk}$ be a field of characteristic 0.\nBy $\\mathsf{Vec}(\\Bbbk)$ and $\\mathsf{vec}(\\Bbbk)$ we denote the categories of $\\Bbbk$-vector spaces and finite-dimensional $\\Bbbk$-vector spaces, respectively.\nFor a $\\Bbbk$-algebra $\\Lambda$, denote by $\\mathsf{J}(\\Lambda)$ the Jacobson radical of $\\Lambda$.\nWe assume $\\mathcal{C}$ is a $\\Bbbk$-linear category.\n\nRecall that a category $\\mathcal{C}$ is called \\emph{Krull--Remak--Schmidt--Azumaya} if any object $X$ is isomorphic to a arbitrary sum $\\bigoplus X_i$, where each $\\End_\\mathcal{C} X_i$ is a local ring, which itself is unique up to isomorphism.\nIf every object is instead isomorphic to a finite sum $\\bigoplus_{i=1}^n X_i$ as above, we say $\\mathcal{C}$ is \\emph{Krull--Remak--Schmidt}.\n\nFinally, we consider (discrete) quivers, continuous generalizations of such quivers, and combinations of the two.\nWhen we draw an arrow, we use a thin line with an arrow head at the end to indicate the direction.\nWhen we draw a continuous line segment, we use a bold line with the arrow head in the middle to indicate the direction;\nsee \\Cref{fig:arrow and line}.\n\n\\begin{figure}[h]\n \\centering\n \\begin{tikzpicture}[decoration={markings, mark=at position 0.5 with {\\arrow{>}}}]\n \\node (left-start) at (0,0) {$\\bullet$};\n \\node (left-end) at (2,0) {$\\bullet$};\n \\node (right-start) at (3,0) {$\\bullet$};\n \\node (right-end) at (5,0) {$\\bullet$};\n \\draw[->] (left-start)--(left-end);\n \n \\draw[very thick, postaction=decorate] (right-start.center)--(right-end.center);\n \n \\end{tikzpicture}\n \\caption{On the left, how we draw arrows. On the right, how we draw line segments.}\n \\label{fig:arrow and line}\n\\end{figure}\n\n\n\\section{Definition and General Results}\\label{sec:definition and general}\n\n\\subsection{$\\Bbbk$-linear categorization}\n\n\\begin{definition}\\label{def:categorification}\n Let $Q$ be a (finite) quiver and $\\Bbbk Q$ its path algebra.\n Let $\\mathcal{Q}$ be the category whose indecomposable objects are the vertices of $Q$ and morphisms between indecomposables $i$ and $j$ are given by\n \\begin{displaymath}\n \\Hom_{\\mathcal{Q}}(i,j) = e_j \\Bbbk Q e_i.\n \\end{displaymath}\n The objects in $\\mathcal{Q}$ are finite direct sums of the indecomposables (and 0).\n The morphisms in $\\mathcal{Q}$ are given by extending bilinearly.\n We call $\\mathcal{Q}$ the \\emph{$\\Bbbk$-linear categorification} of $Q$.\n\\end{definition}\n\n\\begin{example}\\label{ex:discrete categorification}\nLet $Q$ be the following quiver:\n\n\\begin{center}\n\\begin{tikzpicture}[xscale=1.5, yscale=.5]\n \\node (1) at (0,1) {1};\n \\node (2) at (1,2) {2};\n \\node (3) at (1,0) {3};\n \\node (4) at (2,1) {4};\n \n \\draw[->] (1) -- node[pos=0.6, above]{$\\alpha_1$} (2);\n \\draw[->] (1) -- node[pos=0.6, below]{$\\beta_1$} (3);\n \\draw[->] (2) -- node[pos=0.4, above]{$\\alpha_2$} (4);\n \\draw[->] (3) -- node[pos=0.4, below]{$\\beta_2$} (4);\n\\end{tikzpicture}\n\\end{center}\nThen the $\\Bbbk$-linear categorification $\\mathcal{Q}$ is a category with indecomposable objects $1, 2, 3$ and $4$. The morphisms in $\\mathcal{Q}$ are given by paths in $Q$, so for example we have $\\Hom(1,4) \\cong\\Bbbk^2$, while $\\Hom(4,1) = 0$\n\\end{example}\n\n\\begin{proposition}\\label{lem:correspondence between paths and morphisms}\n There is a bijection between nonzero elements in $\\Bbbk Q$ and nonzero morphisms in $\\mathcal{Q}$.\n\\end{proposition}\n\\begin{proof}\n A non-zero element in $\\Bbbk Q$ is a finite sum of paths in $Q$. We can therefore define a map $F$ from the elements in $\\Bbbk Q$ to morphisms in $\\mathcal{Q}$ by specifying the action of $F$ on paths in $Q$. We let this mapping be determined by $\\Hom_{\\mathcal{Q}}(i,j) = e_j \\Bbbk Q e_i$. This map is a bijection by bilinearity of $\\mathcal{Q}$.\n\\end{proof}\n\n\n\\begin{lemma}\\label{prop:same representations}\n Let $\\mathsf{Mod}(\\mathcal{Q})$ be the category of functors $\\mathcal{Q}\\to \\mathsf{Vec}(\\Bbbk)$.\n Then there exists an isomorphism of categories $\\Phi:\\mathsf{Mod}(\\mathcal{Q}) \\to \\mathsf{Rep}(Q)$.\n\\end{lemma}\n\n\\begin{proof}\n Let $F$ be a functor in $\\mathsf{Mod}(\\mathcal{Q})$.\n We now define the corresponding representation $V=\\Phi(F)$.\n Let $M$ be the representation of $Q$ over $\\Bbbk$ where $V(i)=F(i)$ for each $i\\in Q_0$.\n For a path $\\rho$ in $Q$, let $V(\\rho)$ be the $\\Bbbk$-linear map $F(\\rho)$.\n \n Let $f:F\\to G$ be a morphism in $\\mathsf{Mod}(\\mathcal{Q})$.\n Then $\\Phi(f):\\Phi(F)\\to \\Phi(G)$ is defined by the $f_i:F(i)\\to G(i)$ for each $i\\in Q_0$.\n Straightforward computations show that $\\Phi$ respects composition and so it is a functor.\n \n Define $\\Phi^{-1}:\\mathsf{Rep}(Q)\\to \\mathsf{Mod}(\\mathcal{Q})$ in the following way.\n For a representation $V$ of $Q$, let $F=\\Phi^{-1}(V)$ be determined by $F(i)=V(i)$, for each $i\\in Q_0$, and $F(\\rho)=V(\\rho)$ for each path in $Q$.\n Morphisms are defined similarly.\n One may check $\\Phi^{-1}\\Phi$ and $\\Phi\\Phi^{-1}$ are the identity functors on $\\mathsf{Mod}(\\mathcal{Q})$ and $\\mathsf{Rep}(Q)$, respectively.\n\\end{proof}\n\n\\subsection{The Jacobson Radical}\\label{sec:jacobson radical}\n\n\\begin{definition}\\label{def:radical of C}\n Let $\\mathcal{C}$ be a category.\n The \\emph{radical} $\\mathsf{Rad}(\\mathcal{C})$ of $\\mathcal{C}$ is the ideal consisting of\n \\begin{displaymath}\n \\mathsf{Rad}_{\\mathcal{C}}(X,Y) := \\left\\{ f\\in\\Hom_{\\mathcal{C}}(X,Y) \\mid \\forall g\\in\\Hom_{\\mathcal{C}}(Y,X),\\, f\\circ g\\in\\mathsf{J}(\\End_{\\mathcal{C}}(Y)) \\right\\},\n \\end{displaymath}\n for each pair of objects $X,Y$ in $\\mathcal{C}$.\n\\end{definition}\n\n\\begin{proposition}[\\cite{Krause}]\\label{prop:Krause}\n Let $X$ and $Y$ be objects in $\\mathcal{C}$.\n Then $\\mathsf{Rad}_{\\mathcal{C}}(X,Y)=\\mathsf{J}(\\Hom_{\\mathcal{C}}(X,Y))$.\n\\end{proposition}\n\n\\begin{proposition}\\label{prop:easy radical}\n Let $f:X\\to Y$ be an morphism for indecomposable objects $X,Y$ in $\\mathcal{C}$.\n Then $f\\in\\mathsf{Rad}(\\mathcal{C})$ if and only if $f$ is not an isomorphism.\n\\end{proposition}\n\\begin{proof}\n Suppose $f$ is not an isomorphism.\n If $\\mathcal{C}$ does not have cycles we are done.\n If $\\mathcal{C}$ has cycles, let $g:Y\\to X$ be a nonzero morphism.\n Then $f\\circ g\\in\\mathsf{J}(\\End_{\\mathcal{C}}(Y))$ and so $f\\in \\mathsf{Rad}_{\\mathcal{C}}(X,Y)$.\n Reversing the argument shows that if $f\\in\\mathsf{Rad}_{\\mathcal{C}}(X,Y)$ then $f$ is not an isomorphism.\n\\end{proof}\n\nRecall that $\\mathcal{C}$ is semi-simple if every object in $\\mathcal{C}$ is a finite direct sum of simple objects and all such direct sums exist.\n\n\\begin{proposition}\\label{prop:C is basic}\n If $\\mathcal{C}$ is Krull--Remak--Schmidt, then $\\mathcal{C}\/ \\mathsf{Rad}(\\mathcal{C})$ is semi-simple.\n\\end{proposition}\n\\begin{proof}\n Let $\\mathcal{C}$ be a Krull--Remak--Schmidt category.\n Let $X$ and $Y$ be indecomposables in $\\mathcal{C}$ such that $X\\not\\cong Y$.\n Then $\\Hom_{\\mathcal{C}}(X,Y)=\\mathsf{Rad}_{\\mathcal{C}}(X,Y)$ and so $\\Hom_{\\mathcal{C} \/ \\mathsf{Rad}(\\mathcal{C})}(X,Y)=0$.\n Extending bilinearly we see $\\mathcal{C} \/ \\mathsf{Rad}(\\mathcal{C})$ is semi-simple.\n\\end{proof}\n\n\\begin{remark}\\label{rmk:categorification is basic}\n It follows immediately from \\Cref{prop:C is basic} that if $Q$ is a finite acyclic quiver then $\\mathcal{Q} \/\\mathsf{Rad}(\\mathcal{Q})$ is semi-simple.\n\\end{remark}\n\n\\subsection{Admissible Ideals}\\label{sec:admissible ideals}\n\n\\begin{definition}[\\cite{Krause}]\n Let $\\{\\mathcal{C}_i\\}_{i\\in I}$ be a family of full additive subcategories of $\\mathcal{C}$. We have an \\emph{orthogonal decomposition} $\\coprod_{i\\in I} \\mathcal{C}_i$ of $\\mathcal{C}$ if every object $X$ in $\\mathcal{C}$ is isomorphic to a direct sum $\\bigoplus_{i\\in I} X_i$, where $X_i$ is an object of $\\mathcal{C}_i$, and for $X_i\\in \\mathcal{C}_i,X_j\\in C_j$ we have $\\Hom_{\\mathcal{C}}(X_i,X_j)=0$ when $i\\neq j$.\n \n We say $\\mathcal{C}$ is \\emph{connected} if the only orthogonal decomposition of $\\mathcal{C}$ is the trivial one. \n\\end{definition}\n\nAn \\emph{ideal} $\\mathcal{I}$ of a category $C$ is a collection of morphisms is $\\mathcal{C}$ such that for any $f\\in \\mathcal{I}$ and for any $g$ and $h$ such that the composition $gfh $ is defined, the composition $gfh\\in \\mathcal{I}$.\nFor an ideal $\\mathcal{I}$ of $\\mathcal{C}$, we denote by $\\mathcal{I}(X,Y)$ the morphisms in $\\Hom_{\\mathcal{C}}(X,Y)\\cap \\mathcal{I}$.\n\n\\begin{remark}\n For an ideal $\\mathcal{I}$ of $\\mathcal{C}$, the category $\\mathcal{C} \/\\mathcal{I}$ has the same objects as $\\mathcal{C}$.\n The morphisms of $\\mathcal{C} \/ \\mathcal{I}$ are given by $\\Hom_{\\mathcal{C}}(X,Y) \/ \\mathcal{I}(X,Y)$.\n A representation $V: \\mathcal{C} \/ \\mathcal{I} \\to \\Bbbk\\text{vec}$ is also a representation of $\\mathcal{C}$ by precomposition with the quotient functor $\\pi$. Thus we obtain a representation $\\widetilde{V}:\\mathcal{C} \\stackrel{\\pi}{\\to} \\mathcal{C} \/ \\mathcal{I} \\stackrel{V}{\\to} \\Bbbk\\text{vec}$.\n Hence\n we may consider the representations of $\\mathcal{C}\/\\mathcal{I}$ as a subcategory of the representations of $\\mathcal{C}$.\n In particular, representations of $\\mathcal{C}\/\\mathcal{I}$ are those representations $V$ of $\\mathcal{C}$ such that if $f\\in\\mathcal{I}$ then $V(f)=0$.\n\\end{remark}\n\n\\begin{definition}\\label{def:admissible}\n Let $\\mathcal{C}$ be a $\\Bbbk$-linear, Krull--Remak--Schmidt--Azumaya category and $\\mathcal{I}$ an ideal in $\\mathcal{C}$.\n We say $\\mathcal{I}$ is \\emph{admissible} if the following are satisfied.\n \\begin{enumerate}\n \\item For each $f$ in $\\mathcal{I}$, there exists a finite collection of morphisms $g_1,\\ldots,g_n$ not in $\\mathcal{I}$ such that $f = g_n\\circ\\cdots\\circ g_1$.\n \\item For each indecomposable $X$ in $\\mathcal{C}$, the endomorphism ring $\\End_{\\mathcal{C}}(X) \/ \\mathcal{I}(X,X)$ is finite-dimensional.\n \\end{enumerate}\n\\end{definition}\n\nWe remark that, in \\Cref{def:admissible}(2), we do not want to require that there is some $n$ that works for all nonisomorphism endomorphisms $f$.\nSee \\Cref{ex: cycles length} for an explicit example why.\n\n\n\\begin{lemma}\\label{lem: admissible in radical}\n\tLet $\\mathcal{C}$ be a Krull--Remak--Schmidt category and $\\mathcal{I}$ an ideal in $\\mathcal{C}$.\n If $\\mathcal{I}$ is admissible, then $\\mathcal{I}$ is contained in the radical of $\\mathcal{C}$.\n\\end{lemma}\n\n\\begin{proof}\n\tLet $f:X\\rightarrow Y$ be a morphism in $\\mathcal{I}$ between indecomposable objects. Then we know by \\Cref{prop:easy radical} that if $f$ is not contained in the radical, it is an isomorphism. However, if $f$ is an isomorphism, we have $1_X\\in\\mathcal{I}$.\n\tThen \\emph{every} morphism to \/ from $X$ is in $\\mathcal{I}$.\n\tThus, if $1_x=g_n\\circ\\cdots\\circ g_1$ for any composition, both $g_n,g_1\\in \\mathcal{I}$, which contradicts condition \\Cref{def:admissible}(1).\n\tHence $\\mathcal{I}(X,Y)\\subseteq \\mathsf{Rad}(X,Y)$ for indecomposable $X,Y$. \n\t\n\tNow let $X=\\bigoplus_{i=0}^m X_i$ and $Y=\\bigoplus_{j=0}^n Y_j$, where each $X_i$ and $Y_j$ is indecomposable. Consider $f\\in \\mathcal{I}(X,Y)$. We can rewrite $f$ as $f=(f_{ij})$, where $f_{ij}:X_i\\rightarrow Y_j$. By composition with the canonical injections and projections, we see that $f_{ij}\\in \\mathcal{I}(X_i,Y_j)$, so by the above, $f_{ij}\\in \\mathsf{Rad}(X_i,Y_j)$.\n\tThen by linearity, $f\\in \\mathsf{Rad}(X,Y)$.\n\\end{proof}\n\n\nLet $Q$ be a finite quiver and let $\\mathcal{Q}$ be its $\\Bbbk$-linear categorification. Suppose $I$ is an ideal of the path algebra $\\Bbbk Q$. We show how to build an ideal $\\mathcal{I}$ in $\\mathcal{Q}$ from $I$. \n\nFrom the definition of the $\\Bbbk$-linear categorification, we know that each path in $I$ corresponds to a non-zero morphism in $\\mathcal{Q}$, see \\Cref{lem:correspondence between paths and morphisms}. We (na\\\"ively) let $\\mathcal{I}$ be the set of morphisms obtained by mapping $I$ to $\\mathcal{Q}$. We now show that $\\mathcal{I}$ is an ideal of the category $\\mathcal{Q}$. \n\nBy $\\Bbbk$-linearity of $Q$, it is enough to consider morphisms between indecomposable objects.\nSuppose $f\\in\\mathcal{I}(i,j)$ for some $i,j\\in Q_0=\\operatorname{Ind}\\mathcal{Q}$, and let $g:j\\rightarrow k$ and $h:l\\rightarrow i$ be two nonzero morphisms in $\\mathcal{Q}$.\nBy \\Cref{lem:correspondence between paths and morphisms} we know that $f$ corresponds to an element $\\rho$ in $e_j\\Bbbk Qe_i$.\nFurther, $g$ corresponds to an element $\\psi$ in $\\Bbbk Qe_j$ and $h$ corresponds to an element $\\phi$ in $e_i\\Bbbk Q$.\nEach of $\\rho$, $\\phi$, and $\\psi$ are are sums of paths in $Q$ from the respective source and to the respective target.\nWithout loss of generality, due to $\\Bbbk$-linearity, suppose each of $\\rho$, $\\phi$, and $\\psi$ is a path in $Q$.\nWe see $\\psi\\rho\\phi$ is an element of $I$ since $I$ is a two-sided ideal containing $\\rho$.\nThe image of the composition $\\psi\\rho\\phi$ is the composition $gfh$, which must therefore be in $\\mathcal{I}$.\n\n\\begin{proposition}\\label{prop:admissible is correct}\n Let $Q$ be a finite quiver and $I$ an ideal of $\\Bbbk Q$ as a path algebra.\n Let $\\mathcal{Q}$ be the $\\Bbbk$-linear category induced by $Q$ and let $\\mathcal{I}$ be the ideal induced by $I$ in $\\mathcal{Q}$.\n Then $\\mathcal{I}$ is an admissible ideal of $\\mathcal{Q}$ as in \\Cref{def:admissible} if and only if $I$ is an admissible ideal of $\\Bbbk Q$.\n\\end{proposition}\n\n\\begin{proof}\n Let $I$ be an admissible ideal of $\\Bbbk Q$ and $\\rho\\in I$.\n We first prove $\\mathcal I$ satisfies property (1) of \\Cref{def:admissible}.\n Without loss of generality, assume $\\rho$ is a path in $Q$.\n Then $\\rho = \\alpha_n \\alpha_{n-1}\\cdots \\alpha_2\\alpha_1$, where each $\\alpha_i$ is an arrow in $Q$.\n Now, let $f$ be the morphism in $\\mathcal{Q}$ corresponding to $\\rho$ and $g_i$ the morphism in $\\mathcal{Q}$ corresponding to $\\alpha_i$, for each $i$.\n Then we know each $g_i\\notin \\mathcal I$ and have satisfied property (1) of \\Cref{def:admissible}.\n Reversing the argument proves the converse.\n \n If $Q$ has no cycles then the proposition immediately holds for property (2) of \\Cref{def:admissible}.\n So, suppose $Q$ has at least one oriented cycle.\n Since $I \\subset \\mathsf{Rad}^n(\\Bbbk Q)$, for some $n\\geq 2$, we see that $\\mathcal I$ must immediately satisfy property (2) of \\Cref{def:admissible}.\n \n Now suppose $\\mathcal{I}$ satisfies \\Cref{def:admissible}(2).\n Since $Q$ is finite, there are finitely many cycles.\n For each cycle $\\rho$ at each vertex $i$, let \n \\[ m_\\rho= \\min_m \\{ m \\mid \\rho^m \\in \\mathcal{I}(i,i)\\}.\\]\n We know such an $m_\\rho$ exists since $\\End_{\\mathcal{Q}}(i) \/ \\mathcal{I}(i,i)$ is finite-dimensional.\n Let $n_\\rho$ be $m_\\rho$ time the length of $\\rho$.\n Then let \\[ N = \\max\\left(\\max_\\rho \\{n_\\rho\\}\\cup \\{\\text{length of longest path without cycles in }Q\\}\\right).\\]\n Thus, $\\mathsf{Rad}^N(\\Bbbk Q)\\supset I$.\n This concludes the proof.\n\\end{proof}\n\n\\begin{remark}\nThe second half of the proof above can be extended to more general quivers.\nSuppose $\\mathcal{Q}$ is the $\\Bbbk$-linear categorification of a (not necessarily finite) quiver $Q$ with finitely many cycles. Suppose that for each cycle $\\rho$ in the quiver with corresponding morphism $f_\\rho:X\\rightarrow X$, there is some $n\\geq 2$ such that $f_\\rho\\in \\mathcal{I}(X,X)$.\nThen $\\mathcal{I}$ satisfies criterion (2) in \\Cref{def:admissible}. \n\nFor the majority of our examples, this will be the criterion we actually use.\n\\end{remark}\n\n\\begin{example}\\label{ex:discrete ideal}\n Consider the quiver from \\Cref{ex:discrete categorification}. Let $I$ be the commutative ideal generated by $\\{\\alpha_2\\alpha_1-\\beta_2\\beta_1\\}$.\n \n In the $\\Bbbk$-linear categorification, the relation $\\alpha_2\\alpha_1-\\beta_2\\beta_1$ can be written as $\\left [ \\begin{smallmatrix}\\alpha_2 & \\beta_2\\end{smallmatrix}\\right]\\left [ \\begin{smallmatrix}\\alpha_1 \\\\ -\\beta_1\\end{smallmatrix}\\right]$. The ideal generated by this morphism fulfills the criteria for being an admissible ideal. \n\\end{example}\n\n\\subsection{General Results}\\label{sec:general results}\n\n\\begin{proposition}\\label{prop:connected}\n Let $\\mathcal{C}$ be connected, and let $\\mathcal{I}$ be an admissible ideal of morphisms. Then $\\mathcal{C} \/ \\mathcal{I}$ is connected.\n\\end{proposition}\n\n\\begin{proof}\n Assume towards a contradiction that $\\mathcal{C} \/ \\mathcal{I}$ is not connected; then there exists a decomposition of $\\mathcal{C} \/ \\mathcal{I}$ into mutually orthogonal subcategories $\\mathcal{C}'_1, \\cdots, \\mathcal{C}'_n$. We can lift these subcategories to subcategories $\\mathcal{C}_1, \\cdots, \\mathcal{C}_n$ of $\\mathcal{C}$. As $\\mathcal{C}$ is connected, these subcategories cannot all be mutually orthogonal, so assume that there exists some morphism $f:X_i\\rightarrow X_j$, with $X_i\\in \\mathcal{C}_i, X_j\\in \\mathcal{C}_j$ and $i\\neq j$. To preserve mutual orthogonality in $\\mathcal{C}\/\\mathcal{I}$, we must have $f\\in \\mathcal{I}$. Then since $I$ is an admissible ideal, we can write $f=g_m\\circ \\cdots \\circ g_1$, where each of the $g_1, \\cdots, g_m$ are not in $\\mathcal{I}$.\n \n Consider $g_1: X_i\\rightarrow Y$ and denote its image in $\\mathcal{C} \/ \\mathcal{I}$ by $\\overline{g_1}$. As $\\mathcal{C} \/ \\mathcal{I}$ is not connected, we can write $Y=\\bigoplus_{i=1}^n Y_i$, with $Y_i\\in \\mathcal{C}'_i$ and $\\overline{g_1}=(g^1_1,\\cdots g_1^n)$. Now, as $\\overline{g_1}$ is nonzero, $g_1^k$ is nonzero for some $k$. If $k\\neq i$, we have reached a contradiction. If $k=i$, we can repeat the argument with $\\overline{g_2}|_{Y_i}$, eventually reaching a contradiction. \n\\end{proof}\n\n\\begin{proposition}\\label{prop:basic}\n Let $\\mathcal{C}$ be Krull--Remak--Schmidt--Azumaya and $\\mathcal{I}$ an admissible ideal.\n Then $\\mathcal{C} \/ \\mathcal I$ is Krull--Remak--Schmidt--Azumaya.\n\\end{proposition}\n\\begin{proof}\nLet $X$ be an object in $\\mathcal{C} \/ \\mathcal I$. Then $X$ is an object in $\\mathcal{C}$, which we assume to be Krull--Remak--Schmidt--Azumaya. It follows that $X=\\bigoplus_{\\alpha} X_\\alpha$, where $\\End_\\mathcal{C}(X_\\alpha)$ is local. If we can show that $\\End_{\\mathcal{C}\/\\mathcal{I}}(X_\\alpha)$ is local, we are done.\n\nWe know that $\\End_{\\mathcal{C}\/\\mathcal{I}}(X_{\\alpha})=\\End_\\mathcal{C}(X_{\\alpha})\/\\mathcal{I}(X_{\\alpha},X_{\\alpha})$. By property 1 of \\Cref{def:admissible}, the identity on $X_{\\alpha}$ cannot be an element of $\\mathcal{I}(X_{\\alpha},X_{\\alpha})$, so $\\End_{\\mathcal{C}\/\\mathcal{I}}(X_{\\alpha})$ is a nonzero quotient ring of a local ring, which is local by the ideal correspondence theorem for quotient rings.\n\\end{proof}\n\n\\begin{proposition}\\label{prop:radical commutes with admissible}\n Let $\\mathcal{C}$ be a category such that all endomorphism rings are artinian and $\\mathcal{I}$ an admissible ideal.\n Then \\[\\mathsf{Rad}(\\mathcal{C} \/ \\mathcal{I})=\\mathsf{Rad}(\\mathcal{C}) \/ \\mathcal{I}.\\] \n\\end{proposition}\n\\begin{proof}\nThe equation $\\mathsf{Rad}(\\mathcal{C} \/ \\mathcal{I}) = \\mathsf{Rad}(\\mathcal{C}) \/ \\mathcal{I}$ holds if and only if $\\mathsf{Rad}_{\\mathcal{C} \/ \\mathcal{I}}(X,Y) = \\mathsf{Rad}_\\mathcal{C}(X,Y) \/ \\mathcal{I}$ holds for all pairs of objects $X,Y\\in \\mathcal{C}$.\n\nFirst note that if $\\End_{\\mathcal{C} }(Y)$ is artinian, then\n\\begin{align*}\n\t \\mathsf{Rad}_\\mathcal{C}(Y,Y) \/ \\mathcal{I} &= \\mathsf{J}(\\End_{\\mathcal{C} }(Y))\/ \\mathcal{I}= \\mathsf{J}(\\End_{\\mathcal{C} }(Y)+\\mathcal{I})\/\\mathcal{I} \\\\ \n\t &= \\mathsf{J}(\\End_{\\mathcal{C} \/ \\mathcal{I}}(Y)) =\\mathsf{Rad}_{\\mathcal{C} \/ \\mathcal{I}}(Y,Y).\n\\end{align*}\nNow we can see that \n\\begin{align*}\n\tf \\in \\mathsf{Rad}_\\mathcal{C}(X,Y) \/ \\mathcal{I} &\\Leftrightarrow f = f'+\\mathcal{I}(X,Y)\\text{, with } f'\\in \\mathsf{Rad}_\\mathcal{C}(X,Y) \\\\\n\t& \\Leftrightarrow f = f'+\\mathcal{I}(X,Y)\\text{, s.t.\\ } f'\\circ g'\\in \\mathsf{J}(\\End_{\\mathcal{C} }(Y))\\, \\forall\\, g'\\in \\Hom_\\mathcal{C} (Y,X)\\\\\n\t& \\Leftrightarrow f\\circ (g'+\\mathcal{I}(Y,X))\\in \\mathsf{J}(\\End_{\\mathcal{C} }(Y) )\/\\mathcal{I}\\, \\forall \\, g'\\in \\Hom_\\mathcal{C} (Y,X)\\\\\n\t& \\Leftrightarrow f\\circ g \\in \\mathsf{J}(\\End_{\\mathcal{C} \/\\mathcal{I}}(Y) )\\, \\forall\\, g\\in \\Hom_{\\mathcal{C}\/\\mathcal{I}} (Y,X)\\\\\n\t& \\Leftrightarrow f \\in \\mathsf{Rad}_{\\mathcal{C}\/\\mathcal{I}}(X,Y). \n\\end{align*}\n\\end{proof}\nIn particular, if $\\mathcal{C}$ is $\\Bbbk$-linear and Krull--Remak--Schmidt, then all endomorphism rings are artinian and the above Proposition holds.\n\nThe following lemma is useful for our examples in \\Cref{sec:examples}.\n\\begin{lemma}\\label{lem:stack admissible}\n Let $\\mathcal{C}$ be a small, $\\Bbbk$-linear, Krull--Remak--Schmidt--Azumaya category and $\\mathcal{I}$ an admissible ideal in $\\mathcal{C}$.\n Let $\\mathcal{J}$ be an admissible ideal in $\\mathcal{C}\/\\mathcal{I}$ and $\\widetilde{\\Jay}$ the set of morphisms $f$ in $\\mathcal{C}$ such that $f+\\mathcal{I}\\in\\mathcal{J}$ in $\\mathcal{C}\/\\mathcal{I}$. Then the following hold:\n \\begin{enumerate}\n \\item $\\widetilde{\\Jay}$ contains $\\mathcal{I}$.\n \\item $\\widetilde{\\Jay}$ is an ideal.\n \\item $\\widetilde{\\Jay}$ is admissible in $\\mathcal{C}$.\n \\item $(\\mathcal{C}\/\\mathcal{I})\/\\mathcal{J} \\simeq \\mathcal{C}\/\\widetilde{\\Jay}$.\n \\end{enumerate} \n\\end{lemma}\n\\begin{proof}\n \\textbf{1.} Let $f\\in\\mathcal{I}$.\n Then $f\\mapsto 0$ in $\\mathcal{C}\/\\mathcal{I}$.\n Since all zero morphisms in $\\mathcal{C}\/\\mathcal{I}$ are in $\\mathcal{J}$, we see $f\\in \\widetilde{\\Jay}$.\n \n \\textbf{2.} Let $f:X\\to Y$ be nonzero in $\\widetilde{\\Jay}$ and let $g:Y\\to Z$ be nonzero in $\\mathcal{C}$.\n Then $f+\\mathcal{I}$ and $g+\\mathcal{I}$ are in $\\mathsf{Mor}(\\mathcal{C}\/\\mathcal{I})$.\n So, $(g+\\mathcal{I})\\circ(f+\\mathcal{I})$ is in $\\mathcal{J}$ and is equal to $gf + \\mathcal{I}$.\n Then $gf \\in \\widetilde{\\Jay}$.\n \n \\textbf{3.} Let $f\\in\\widetilde{\\Jay}$.\n Then $f+\\mathcal{I}\\in\\mathcal{J}$ and, by assumption, there exists $(g_1+\\mathcal{I}),\\ldots,(g_n+\\mathcal{I})$ in $\\mathsf{Mor}(\\mathcal{C}\/\\mathcal{I})\\setminus\\mathcal{J}$ such that \\[ f + \\mathcal{I} = (g_n+\\mathcal{I})\\circ (g_{n-1}+\\mathcal{I})\\circ\\cdots\\circ (g_2+\\mathcal{I})\\circ(g_1+\\mathcal{I}).\\]\n Then for each $g_i$ there is $h_i\\in\\mathcal{I}$ such that $g_i+h_i\\mapsto g_i+\\mathcal{I}$ and \\[f = (g_n+h_n)\\circ(g_{n-1}+h_{n-1})\\circ\\cdots\\circ(g_2+h_2)\\circ (g_1+h_1). \\]\n Since each $g_i\\notin \\mathcal{J}$, we know each $g_i+h_i\\notin\\widetilde{\\Jay}$ and so $f$ is a finite composition of morphisms not in $\\widetilde{\\Jay}$.\n Additionally, for any nonzero, nonisomorphism endomorphism $f$, we have $f^n\\in\\mathcal{I}$ for some $n\\in\\mathbb{N}$.\n Then $f^n\\in\\widetilde{\\Jay}$ by statement 1.\n Therefore, $\\widetilde{\\Jay}$ is admissible.\n \n \\textbf{4.} Recall $\\mathsf{Ob}((\\mathcal{C}\/\\mathcal{I})\/\\mathcal{J}) = \\mathsf{Ob}(\\mathcal{C}\/\\widetilde{\\Jay})$.\n We now produce a bijection between $\\mathsf{Mor}((\\mathcal{C}\/\\mathcal{I})\/\\mathcal{J})$ and $\\mathsf{Mor}(\\mathcal{C}\/\\widetilde{\\Jay})$ by producing bijections \\[ \\phi_{X,Y}:\\Hom_{\\mathcal{C}\/\\widetilde{\\Jay}}(X,Y) \\to \\Hom_{(\\mathcal{C}\/\\mathcal{I})\/\\mathcal{J}}(X,Y) \\] for each ordered pair $X,Y$ of objects.\n \n Let $f+\\widetilde{\\Jay}\\in\\Hom_{\\mathcal{C}\/\\widetilde{\\Jay}}(X,Y)$.\n Then there exists $g\\in\\widetilde{\\Jay}\\subset \\mathsf{Mor}(\\mathcal{C})$ such that $f+g\\mapsto f+\\widetilde{\\Jay}\\in\\mathsf{Mor}(\\mathcal{C}\/\\widetilde{\\Jay})$.\n If $g\\in\\mathcal{I}$ then $f+g\\mapsto f+\\mathcal{I}\\in\\mathsf{Mor}(\\mathcal{C}\/\\mathcal{I})$; otherewise $f+g\\mapsto f+g+\\mathcal{I}\\in\\mathsf{Mor}(\\mathcal{C}\/\\mathcal{I})$.\n In either case, $f+g\\mapsto f+\\mathcal{J}$ in $\\Hom_{(\\mathcal{C}\/\\mathcal{I})\/\\mathsf{J}}(X,Y)$.\n We define $\\phi_{X,Y}(f+\\widetilde{\\Jay}):= f+\\mathcal{J}$.\n \n It is immediate that $\\phi_{X,Y}$ is injective.\n Suppose $f+\\mathcal{J}\\in\\Hom_{(\\mathcal{C}\/\\mathcal{I})\/\\mathcal{J}}$.\n Then there exists $g+\\mathcal{I}$ in $\\Hom_{\\mathcal{C}\/\\mathcal{I}}(X,Y)$ such that $f+g+\\mathcal{I}\\mapsto f+\\mathcal{J}$.\n Then there exists $h\\in\\Hom_{\\mathcal{C}}(X,Y)$ such that $f+g+h\\mapsto f+g+\\mathcal{I}$.\n But this means $g+h\\in\\widetilde{\\Jay}$ and so $f+(g+h)\\mapsto f+\\widetilde{\\Jay}$ in $\\Hom_{\\mathcal{C}\/\\widetilde{\\Jay}}(X,Y)$.\n Thus, $\\phi_{X,Y}$ is surjective.\n\\end{proof}\n\n\\section{Relations}\\label{sec:relations}\n\nIn this section we look at two types of admissible ideals: those generated by point relations (\\Cref{subsec:point relations}) and those generated by length relations (\\Cref{subsec:length}).\nThese generalize a relation generated by a single path of length two and relations generated by all paths of a particular length, respectively.\n\n\\subsection{Point Relations}\\label{subsec:point relations}\nHere we generalize relations generated by a single path of length 2 to a point relation (\\Cref{def:point relation}) and prove this generates an admissible ideal (\\Cref{thm:point relations are admissible}).\nWe then give examples that point to a continuous version of a gentle algebra (\\Cref{ex:monomial points,ex:crossing real lines}).\n\n\\begin{definition}\\label{def:decomposition point}\n Let $f:X\\to Y$ be a nonisomorphism between indecomposables in $\\mathcal{C}$.\n A \\emph{decomposition point} $Z$ in $\\mathcal{C}$ is an indecomposable object such that there exists nonisomorphisms $g:X\\to Z$ and $h:Z\\to Y$ where $f=h\\circ g$.\n\\end{definition}\n\n\\begin{definition}\\label{def:acyclic morphism}\n Let $f:X\\to Y$ be a nonisomorphism of indecomposables in $\\mathcal{C}$, $Z$ a decomposition point of $f$, and $f=g\\circ h$ such a decomposition.\n We call $f$ an \\emph{acyclic morphism} if for all pairs $g':X\\to Z$ and $h':Z\\to Y$ such that $f=h'\\circ g'$ then $h'$ and $g'$ are scalar multiples of $h$ and $g$, respectively.\n\\end{definition}\n\nNote that an acyclic morphism cannot be irreducible. (I.e., it must be a path of length at least 2 in a quiver.)\n\n\\begin{definition}\\label{def:point relation}\n Let $f$ be an acyclic morphism and $Z$ a decomposition point of $f$.\n Let $P$ be the set of all nonisomorphisms $g$ between indecomposables satisfying the following.\n \\begin{itemize}\n \\item There exists $h_1$ and $h_2$ morphisms of indecomposables such that $f=h_2\\circ g\\circ h_1$.\n \\item We have $Z$ as a decomposition point of $g$.\n \\end{itemize}\n Let $\\mathcal P_{f,Z}$ be the ideal in $\\mathcal{C}$ generated by $P$.\n We call $\\mathcal P_{f,Z}$ the \\emph{point relation through $Z$ by $f$}.\n\\end{definition}\n\n\\begin{definition}\\label{def:admissible point relations}\n Let $\\{\\mathcal P_\\alpha\\}$ be a collection of point relations in $\\mathcal{C}$.\n We say $\\{\\mathcal P_\\alpha\\}$ is \\emph{admissible} if each morphism of indecomposables appears in at most finitely-many $\\mathcal P_\\alpha$.\n\\end{definition}\n\n\\begin{theorem}\\label{thm:point relations are admissible}\n Let $\\{\\mathcal P_\\alpha\\}$ be an admissible collection of point relations in $\\mathcal{C}$ and let $\\mathcal{I}=\\langle \\bigcup_\\alpha P_\\alpha\\rangle$.\n Suppose also that for each indecomposable $X$ in $\\mathcal{C}$, we have $\\End_{\\mathcal{C}}(X) \/ \\mathcal{I}(X,X)$ is finite dimensional.\n Then $\\mathcal{I}$ is an admissible ideal.\n\\end{theorem}\n\\begin{proof}\nWe satisfy \\Cref{def:admissible}(2) by assumption.\n\nNow suppose $f\\in \\mathcal{I}$; we show that $f$ can be written as a finite composition of morphisms not in $\\mathcal{I}$.\nSince $f$ is a finite sum of morphisms of indecomposables, we assume without loss of generality $f$ is a morphism of indecomposables.\nThen, by assumption there are at most finitely-many $\\mathcal P_\\alpha$ such that $f\\in \\mathcal P_\\alpha$.\n\nWe proceed by induction beginning with $f$ is only in $\\mathcal P_1$.\nLet $f_1$, $P_1$, and $Z_1$ be as in \\Cref{def:point relation}.\nThen there exists morphisms $h_1$ and $h_2$ in $Mor(\\mathcal{C})$ and $g\\in P_1$ such that $f = h_2\\circ g \\circ h_1$.\nFurther, $g=h'_2\\circ h'_1$ where the target of $h'_1$ is $Z_\\alpha$ and the source of $h'_2$ is $Z_\\alpha$.\nSo, let $g_1 = h'_1\\circ h_1$ and let $g_2=h_2\\circ h'_2$.\nNote that neither $g_1$ nor $g_2$ is in $\\mathcal P_1$.\nFurther, neither $g_1$ nor $g_2$ is in $\\mathcal{I}$ or else $f$ would be in another $\\mathcal P_\\alpha$ as well.\nThus, we have our desired decomposition.\n\nNow assume that if $f$ is in $n$ of the $\\mathcal P_\\alpha$, then $f$ is a finite composition of morphisms not in $\\mathcal{I}$.\nSuppose $f$ is in $n+1$ of the $\\mathcal P_\\alpha$ and denote one of them by $\\mathcal P_1$.\nLet $f_1$, $P_1$, and $Z_1$ be as before for $\\mathcal P_1$.\nWe find $g_1$ and $g_2$ as before, but they may be in $\\mathcal{I}$.\nHowever, each $g_1$ and $g_2$ may only be in $n$ or fewer $\\mathcal P_\\alpha$ and so are a finite composition of morphisms not in $\\mathcal{I}$.\nTherefore, $f$ is a finite composition of morphisms not in $\\mathcal{I}$.\n\\end{proof}\n\n\\begin{example}[Discrete quiver]\\label{ex:monomial points}\nLet $Q$ be a discrete quiver. Then any quadratic monomial relation in $Q$ corresponds to a point relation in the $\\Bbbk$-linear categorification of $Q$.\n\nIn particular, any gentle algebra can be obtained by considering a quiver with point relations. \n\\end{example}\n\n\\begin{example}[Continuous ``gentle'', crossing real lines]\\label{ex:crossing real lines}\nConsider two copies of the real line, labeled $\\mathbb{R}$ and $\\mathbb{R}'$. We label the numbers in $\\mathbb{R}$ by $x$ and the numbers in $\\mathbb{R}'$ by $x'$. Identify $0$ and $0'$ and label the category of $\\Bbbk$-representations of the resulting partially ordered set by $\\mathcal{C}$. \n\\begin{figure}\n \\centering\n \\begin{tikzpicture}[inner sep = 0cm, outer sep = 0cm, xscale = 1, decoration={\n markings,\n mark=at position 0.5 with {\\arrow{>}}}, very thick\n ]\n \\coordinate (topleft) at (0,1);\n \\coordinate (bottomleft) at (0,0);\n \\coordinate (topright) at (10,1);\n \\coordinate (bottomright) at (10,0);\n \\coordinate (mid) at (5,0.5);\n \\draw[blue, postaction=decorate] (topleft) ..controls +(4,0).. (mid); \n \\draw[red, postaction=decorate] (mid) ..controls +(1,-.5).. (bottomright);\n \\draw[red, postaction=decorate] (bottomleft) ..controls +(4,0).. (mid); \n \\draw[blue, postaction=decorate] (mid) ..controls +(1,.5).. (topright);\n \\draw[fill=white] (mid) circle[radius=1mm];\n \\end{tikzpicture}\n \\caption{The category considered in \\Cref{ex:crossing real lines}. The two copies of the real line have been drawn in different colours}\n \\label{fig:my_label}\n\\end{figure}\n\nLet $\\mathcal P$ be the point relation at $0$ generated morphisms starting in $\\mathbb{R}_{<0}$ and ending in $\\mathbb{R}'_{>0}$. Dually, let $\\mathcal P'$ be the point relation at $0$ generated morphisms starting in $\\mathbb{R}'_{<0}$ and ending in $\\mathbb{R}_{>0}$. The collection $\\{\\mathcal P, \\mathcal P'\\}$ generates an admissible ideal.\nIn later work, we will argue that this $\\mathcal{C}$ with this ideal yields a continuous generalization of a gentle algebra. \n\n\\begin{remark}\nIf we do not assume that $\\End_{\\mathcal{C}}(X) \/ \\mathcal{I}(X,X)$ is finite dimensional in our hypothesis of \\Cref{thm:point relations are admissible}, it is possible that we do not have an admissible ideal.\nSee \\Cref{ex:big wedge}.\n\\end{remark}\n\n\\end{example}\n\\subsection{Length Relations}\\label{subsec:length}\nWe now generalize relations generated by all paths of a certain length to length relations.\nTo do this we define a way of measuring length in our category (\\Cref{def:weakly archimedean monoid,def:category with length in Lambda}) and provide examples (\\Cref{ex:weakly Archimedian monoids,ex:categories with length in Lambda}).\nThen we define the length relations (\\Cref{def:length relation}) and provide examples (\\Cref{ex:length relations}) and prove that length relations generate admissible ideals (\\Cref{thm:length relations are admissible}).\nIn \\Cref{apx:length} we discuss the proof of \\Cref{thm:length relations are admissible} (\\Cref{subsec:need weakly archimedean}), why we require the specific setup that we have (\\Cref{sec:more on length}), and compare our notion of length to the notion of a metric on a category, introduced by Lawvere \\cite{L73} (\\Cref{subsec:length vs metric}).\n\nRecall a commutative monoid $\\Lambda$ is a set with an associative, commutative, binary operation $+_{\\Lambda}:\\Lambda\\times\\Lambda \\to \\Lambda$ and an identity 0.\n\\begin{definition}\\label{def:weakly archimedean monoid}\n Let $\\Lambda$ be a commutative monoid.\n We say $\\Lambda$ is \\emph{weakly Archimedian} if it satisfies the following.\n \\begin{itemize}\n \\item There is a total order $\\leq$ on $\\Lambda$.\n \\item If $\\lambda\\neq 0$ then $\\lambda>0$.\n \\item If $\\lambda_1>\\lambda_2$ then, for any $\\lambda_3$, we have $\\lambda_1+\\lambda_3> \\lambda_2+\\lambda_3$ or $\\lambda_1+\\lambda_3 = \\lambda_2+\\lambda_3=\\max \\Lambda$.\n \\item For all $0<\\lambda_1<\\lambda_2$ in $\\Lambda$, there exists $n\\in\\mathbb{N}$ such that\n \\begin{displaymath}\n n\\lambda_1 := \\underbrace{\\lambda_1 +_{\\Lambda} \\lambda_1 +_{\\Lambda} \\cdots +_{\\Lambda} \\lambda_1}_{n} \\geq \\lambda_2.\n \\end{displaymath}\n \\end{itemize}\n\\end{definition}\n\n\\begin{example}\\label{ex:weakly Archimedian monoids}\n We give three examples, two of which the reader might expect.\n \\begin{enumerate}\n \\item The set $\\mathbb{N}$ with the usual total order and $+_{\\mathbb{N}}$ given in the usual way is weakly Archimedian.\n \\item The set $\\mathbb{R}_{\\geq 0}$ with the usual total order and $+_{\\mathbb{R}}$ given in the usual way is weakly Archimedian.\n \\item\\label{ex:weakly Archimedian monoids:with max} Let $\\Lambda = \\{0,1,2,\\ldots,n-1,n,\\infty\\}$. Let $+_{\\Lambda}$ be given by\n \\begin{displaymath}\n \\lambda_1+_{\\Lambda}\\lambda_2 = \\begin{cases}\n \\lambda_1+_{\\mathbb{N}}\\lambda_2 & (\\lambda_1+_{\\mathbb{N}}\\lambda_2)\\leq n \\\\\n \\infty & \\text{otherwise}.\n \\end{cases}\n \\end{displaymath}\n For the total order, we say $0<1<2<\\cdots0$ and for each $\\lambda<\\ell(f)$ there are $g,h\\in\\mathsf{Mor}(\\widehat{\\mathcal{C}})$ such that $f=h\\circ g$ and either $\\ell(g)=\\lambda$ or $\\ell(h)=\\lambda$.\n \\end{enumerate}\n\\end{definition}\n\n\\begin{example}\\label{ex:categories with length in Lambda}\n We give some existing examples of \\Cref{def:category with length in Lambda}.\n \\begin{enumerate}\n \\item\\label{ex:categories with length in Lambda:classic} Let $Q$ be a quiver, $\\mathcal{Q}$ the $\\Bbbk$-linear categorification of $Q$, and $\\widehat{\\mathcal{Q}}$ a stem category of $\\mathcal{Q}$ whose morphisms are generated by arrows.\n Let $\\Lambda=\\mathbb{N}$ and set $\\ell(\\alpha)=1$ for each morphism in $\\widehat{\\mathcal{Q}}$ from an arrow $\\alpha$ in $Q$.\n Then $\\mathcal{Q}$ has length in $\\mathbb{N}$.\n \\item\\label{ex:categories with length in Lambda:AR} Let $\\mathcal{Q}$ be the additive $\\Bbbk$-linearization of a continuous quiver $\\widehat{\\mathcal{Q}}$ of type $A$ as in \\cite{IRT}.\n Then $\\widehat{\\mathcal{Q}}$ is a stem category of $\\mathcal{Q}$.\n \n Define $\\ell: \\mathsf{Mor}(\\widehat{\\mathcal{Q}})\\to\\mathbb{R}_{\\geq 0}$ by $\\ell(g_{x,y}) = |x-y|$ where $g_{x,y}$ is the unique nonzero morphism in $\\widehat{\\mathcal{Q}}$ from $x$ to $y$.\n Then $\\mathcal{Q}$ has length in $\\mathbb{R}_{\\geq 0}$.\n Note $\\mathcal{Q}$ is also a category with a metric.\n See \\Cref{subsec:length vs metric}.\n \\end{enumerate}\n\\end{example}\n\nWe now define a length relation.\n\\begin{definition}\\label{def:length relation}\n Let $\\Lambda$ be a weakly Archimedian monoid and $\\mathcal{C}$ a category with length in $\\Lambda$ with stem category $\\widehat{\\mathcal{C}}$.\n Consider $\\Lambda_1,\\Lambda_2$ subsets of $\\Lambda$ such that $\\Lambda_1\\amalg\\Lambda_2=\\Lambda$, $|\\Lambda_1|\\geq 2$, and for all $\\lambda_1\\in\\Lambda_1, \\lambda_2\\in\\Lambda_2$ we have $\\lambda_1<\\lambda_2$.\n Then the set $\\ell^{-1}(\\Lambda_2)$ in $\\widehat{\\mathcal{C}}$ generates an ideal $\\mathcal{I}$ in $\\mathcal{C}$.\n We call $\\mathcal{I}$ a \\emph{length relation}.\n\\end{definition}\n\n\\begin{remark}\\label{rmk:length is not a number}\n It is possible that $\\Lambda_1$ has no maximum element and $\\Lambda_2$ has no minimum element.\n (Consider, for example, $\\Lambda=\\mathbb{Q}_{\\geq 0}$.)\n Thus, we may not always be able to say that we are taking ``paths longer than $\\lambda$'' for some $\\lambda\\in\\Lambda$.\n\\end{remark}\n\n\\begin{example}\\label{ex:length relations}\n We give three examples of length relations.\n \\begin{enumerate}\n \\item Let $Q$ be a quiver and $\\mathcal{Q}$ its $\\Bbbk$-linear categorification, which has length in $\\mathbb{N}$ (\\Cref{ex:categories with length in Lambda}(\\ref{ex:categories with length in Lambda:classic})).\n Let $\\widehat{\\mathcal{Q}}$ be the stem category of $\\mathcal{Q}$ seen, effectively, as $Q$ embedded in $\\mathcal{Q}$.\n Let $\\Lambda_1 = \\{0,1,2\\}$ and $\\Lambda_2=\\{3,4,5,\\cdots\\}$.\n Then $\\mathcal{I}=\\langle \\ell^{-1}(\\Lambda_2)\\rangle $ is the set of morphisms in $\\mathcal{Q}$ generated by paths with length $\\geq 3$ in $Q$.\n \\item Any Nakayama algebra where the relations have constant length $l$ can be realized as the $\\Bbbk$-linear categorification of its underlying quiver with length relations of length $l$ in $\\mathbb{N}$.\n \\item\\label{ex:length relations:continuous A} Let $\\mathcal{Q}$ and $\\widehat{\\mathcal{Q}}$ be as in \\Cref{ex:categories with length in Lambda}(\\ref{ex:categories with length in Lambda:AR}).\n Recall $\\mathcal{Q}$ has length in $\\mathbb{R}_{\\geq 0}$.\n Let $\\Lambda_1 =[0,4]$ and $\\Lambda_2=(4,+\\infty)$.\n Then $\\langle\\ell^{-1}(\\Lambda_2)\\rangle$ is the set of morphisms in $\\mathcal{Q}$ of length \\emph{strictly greater than $4$}.\n \\end{enumerate}\n\\end{example}\n\n\\begin{theorem}\\label{thm:length relations are admissible}\n Let $\\Lambda$ be a weakly Archimedian monoid, $\\mathcal{C}$ a category with length in $\\Lambda$ with stem category $\\widehat{\\mathcal{C}}$, and $\\mathcal{I}$ a length relation.\n If $\\End_{\\widehat{\\mathcal{C}}}(X)$ is a finitely-generated monoid, for each $X\\in\\mathsf{Ob}(\\widehat{\\mathcal{C}})$, then $\\mathcal{I}$ is an admissible ideal.\n\\end{theorem}\n\\begin{proof}\n If $\\mathcal{I}=\\emptyset$ then condition (1) is vacously satisfied.\n Assume $\\mathcal{I}\\neq\\emptyset$, let $f\\in \\mathcal{I}$ such that $f\\in\\mathsf{Mor}(\\widehat{\\mathcal{C}})$, and let $\\lambda\\in\\Lambda_1$ such that $\\lambda>0$.\n Then there is $n\\in\\mathbb{N}$ such that $n\\lambda \\geq \\ell(f)$.\n Thus, there is some decomposition $f=g_n\\circ\\cdots\\circ g_1$ where $\\ell(g_i)\\in\\Lambda_1$ for each $g_i$.\n Thus, each $g_i$ is not in $\\mathcal{I}$.\n \n Since $\\End_{\\widehat{\\mathcal{C}}}(X)$ is a finitely-generated monoid, let $m$ be the number of generators and let $\\{f_i\\}_{i=1}^m$ be the set of generators.\n Let\n \\[ N = \\max_i \\{ \\min_n \\{ n\\ell(f_i) \\mid n\\ell(f_i)\\in\\Lambda_2\\}\\}.\\]\n Then\n \\[ \\dim_{\\Bbbk}(\\End_{\\mathcal{C}}(X) \/ \\mathcal{I}(X,X)) \\leq m\\cdot N + 1,\\]\n where we need the ``$+1$'' to account for the identity in $\\End_{\\widehat{\\mathcal{C}}}(X)$.\n Therefore, $\\mathcal{I}$ is an admissible ideal.\n\\end{proof}\n\n\n\n\\section{Examples}\\label{sec:examples}\n\\begin{figure}[b]\n \\centering\n \\begin{subfigure}[b]{0.4\\textwidth}\n \\centering\n \n\\begin{tikzpicture}[xscale=2]\n\\node (1) at (0,0) {1};\n\\node (2) at (1,1) {2};\n\\node (3) at (1,0) {3};\n\\node (4) at (1,-1) {4};\n\\node (5) at (2,0) {5};\n\n\\draw[->] (1) -- node[pos=0.6, above]{$\\alpha_1$} (2);\n\\draw[->] (1) -- node[pos=0.7, above]{$\\beta_1$} (3);\n\\draw[->] (1) -- node[pos=0.6, above]{$\\gamma_1$} (4);\n\\draw[->] (2) -- node[pos=0.4, above]{$\\alpha_2$} (5);\n\\draw[->] (3) -- node[pos=0.3, above]{$\\beta_2$} (5);\n\\draw[->] (4) -- node[pos=0.4, above]{$\\gamma_2$} (5);\n\\end{tikzpicture}\n \\caption{}\n \\label{fig:three-path-discrete}\n \\end{subfigure}\n \\begin{subfigure}[b]{0.4\\textwidth}\n \\centering\n \\begin{tikzpicture}[xscale=2, inner sep = 0cm, outer sep = 0cm, very thick,decoration={\n markings,\n mark=at position 0.5 with {\\arrow{>}}}\n ]\n\\node (start) at (0,0) {$\\bullet$};\n\\node (end) at (2,0) {$\\bullet$};\n\\node at (-.1,0){$0$};\n\\node at (2.1,0){$1$};\n\n\\draw[postaction={decorate}] (start) .. controls (.5,1.3) and (1.5,1.3).. node[above=2pt, pos=0.6]{$\\alpha$} (end);\n\\draw[postaction={decorate}] (start) -- node[pos=0.6, above=2pt]{$\\beta$} (end);\n\\draw[postaction={decorate}] (start) .. controls (.5,-1.3) and (1.5,-1.3).. node[pos=0.6, above=2pt]{$\\gamma$} (end);\n\\end{tikzpicture}\n \\caption{}\n \\label{fig:three-path-continuous}\n \\end{subfigure}\n \\caption{The quivers considered in \\Cref{ex: discrete 3 paths,ex: cts 3 paths}, respectively.}\n \\label{fig:three-path-quivers}\n\\end{figure}\n\n\\begin{example}\\label{ex: discrete 3 paths}\nConsider the (discrete) quiver $Q$ shown in \\Cref{fig:three-path-discrete}.\nThe relation $\\alpha_2\\alpha_1-2\\beta_2\\beta_1+3\\gamma_2\\gamma_1$ generates an admissible ideal in the classical sense. In the $\\Bbbk$-linear categorification of $Q$, we rewrite the relation as the composition of morphisms\n\\[1\\xrightarrow{\\left [ \\begin{smallmatrix}\\alpha_1 \\\\ -2\\beta_1\\\\3\\gamma_1\\end{smallmatrix}\\right]} 2\\oplus 3\\oplus 4\\xrightarrow{\\left [ \\begin{smallmatrix}\\alpha_2 & \\beta_2 &\\gamma_2\\end{smallmatrix}\\right]} 5,\n \\]\n which generates an admissible ideal\n \n\\end{example}\n\n\n\\begin{example} \\label{ex: cts 3 paths}\n\nConsider the continuous analogue of \\Cref{ex: discrete 3 paths}, displayed in \\Cref{fig:three-path-continuous}.\nWe consider a similar relation $\\alpha-2\\beta+3\\gamma$.\nLet $X$, $Y$, and $Z$ be points on the interior of the paths $\\alpha$, $\\beta$, and $\\gamma$, respectively.\nThen $\\alpha=\\alpha_2\\alpha_1$ where $\\alpha_1:0\\to X$ and $\\alpha_2:X\\to 1$.\nWe similarly write $\\beta=\\beta_2\\beta_1$ and $\\gamma=\\gamma_2\\gamma_1$.\nThen the relation $\\alpha-2\\beta+3\\gamma$ can be written as the composition\n\\[0\n\\xrightarrow{\\left [ \\begin{smallmatrix}\\alpha_1 \\\\ -2\\beta_1\\\\3\\gamma_1\\end{smallmatrix}\\right]}\nX\\oplus Y\\oplus Z\n\\xrightarrow{\\left [ \\begin{smallmatrix}\\alpha_2 & \\beta_2 &\\gamma_2\\end{smallmatrix}\\right]}\n1,\n \\]\nand it generates an admissible ideal.\n\\end{example}\n\n\\begin{example}[Real line with point relations on integers]\\label{ex:real line integers}\nLet $\\mathcal{Q}$ be the additive $\\Bbbk$-linearization of $\\mathbb{R}$ as a category where paths move upwards.\nFor a point $c\\in \\mathbb{R}$, let the (unique) point relation at $c$ be $\\mathcal P_c$. The collection of point relations on the integers, $\\{\\mathcal P_z\\}_{z\\in \\mathbb{Z}}$ generates an admissible ideal by \\Cref{thm:point relations are admissible}.\n\nThe ``Auslander--Reiten space'' of the representations of this quiver is shaped like a mountain range; it is a set of triangles joined at their bottom vertices, see \\Cref{fig:AR mountain range}\n\\begin{figure}\n \\centering\n \\begin{subfigure}{\\textwidth}\n \\centering\n \\begin{tikzpicture}\n \\draw[dotted, thick] (-1.5,0) -- (0,0)--(-1.5,1);\n \\draw[dotted,thick] (10.5,1) -- (9,0) -- (10.5,0);\n \\draw[] (0,0) -- (1.5,1)--(3,0)--(4.5,1)--(6,0)--(7.5,1)--(9,0)--(0,0);\n \\end{tikzpicture}\n \\caption{The ``Auslander--Reiten space'' of representations of the quiver from \\Cref{ex:real line integers}}\n \\label{fig:AR mountain range}\n \\end{subfigure}\n \n \\begin{subfigure}{\\textwidth}\n \\begin{tikzpicture}\n \\draw[color=white] (0,0)--(0,1.3);\n \\draw[dotted, thick] (-1.5,0) -- (0,0)--(-1,1)--(-1.5,1);\n \\draw[dotted,thick] (10.5,1) -- (10,1) -- (9,0) -- (10.5,0);\n \\draw[] (0,0) -- (1,1)--(2,1)--(3,0)--(4,1)--(5,1)--(6,0)--(7,1)--(8,1)--(9,0)--(0,0);\n \\end{tikzpicture}\n \n \\caption{The ``Auslander--Reiten space'' for representations of $\\mathbb{R}$ modulo paths longer than $s$ and modulo paths that go through any $x\\in r\\mathbb{Z}$, where $r>s\\in\\mathbb{R}_{>0}$.}\\label{fig:AR chopped mountains}\n \\end{subfigure}\n \\caption{The ``Auslander--Reiten spaces'' of representations of the quivers in \\Cref{ex:real line integers,ex:real line points}}\n \\label{fig:AR spaces}\n\\end{figure}\n\\end{example}\n\n\\begin{example}[Circle with length\/Kupisch relations]\\label{ex:circle length}\n Let $\\mathcal{Q}$ be the additive $\\Bbbk$-linearization of a continuous quiver $\\widehat{\\mathcal{Q}}$ of type $\\widetilde{A}$ as in \\cite{HR}.\n Define $\\ell: \\mathsf{Mor}(\\widehat{\\mathcal{Q}})\\to\\mathbb{R}_{\\geq 0}$ by $\\ell(f)=\\phi-\\theta + 2n\\pi$ where $f:e^{i\\theta}\\to e^{i\\phi}$, and $0\\leq \\phi-\\theta<2\\pi$, and $n$ is the number of full rotations around the circle at $e^{i\\theta}$ before moving to $e^{i\\phi}$.\n Then $\\mathcal{Q}$ has length in $\\mathbb{R}_{\\geq 0}$.\n If $\\mathcal{Q}$ is acyclic, we may replace $\\mathbb{R}_{\\geq 0}$ with $\\Lambda=[0,2\\pi)\\cup\\{\\infty\\}$ and define $+_\\Lambda$ similarly to \\Cref{ex:weakly Archimedian monoids}(\\ref{ex:weakly Archimedian monoids:with max}).\n \n \\begin{figure}\n \\centering\n \\begin{tikzpicture}[very thick, decoration={markings,mark=at position 0.5 with {\\arrow{>}}}]\n \\draw[radius=3cm] (0,0) circle;\n \\draw[radius=2.8cm, color=red,postaction=decorate] (180:2.8cm) arc[start angle =180, delta angle= 180] node[above left, pos=0.666, red]{\\XSolidBrush};\n \\draw[radius=3.2cm, color=blue, postaction=decorate] (0:3.2cm) arc[start angle = 0, delta angle= 90] node[above right, pos=0.666, blue]{\\CheckmarkBold};\n \\foreach \\x in {0, 30,...,360}\n \t\\draw[thin] (\\x:2.9cm) -- (\\x:3.1cm); \n \\end{tikzpicture}\n \\caption{The circle with length relations as described in \\Cref{ex:circle length}. In this figure, the relations have length $\\frac{2\\pi}{3}$.}\n \n \\label{fig:circle length}\n \\end{figure}\n \n Now assume $\\widehat{\\mathcal{Q}}$ has cyclic orientation.\n Let $\\kappa$ be a Kupisch function as in \\cite[Definition~3.9]{RZ}.\n That is, $\\kappa:\\mathbb{R}\\to\\mathbb{R}_{> 0}$ is a function such that $\\kappa(t)+t>t$ and $\\kappa(t+1)=\\kappa(t)$, for all $t\\in\\mathbb{R}$.\n This yields a map $\\mathbb{S}^1\\to \\mathbb{R}_{>0}$ where $\\mathbb{S}^1=[0,1]\/\\{0\\sim 1\\}$.\n If $\\kappa$ is constant with value $a$, then this yields a length relation where $\\Lambda_1=[0,a]$ and $\\Lambda_2=(a,+\\infty)$.\n If $\\kappa$ is not constant, then we do not have a length relation.\n However, if $\\kappa$ does not have any separation points \\cite[Definition~4.2]{RZ}, then $\\kappa$ still induces an admissible ideal.\n\\end{example}\n\n\\begin{example}[Real line with length and point relations]\\label{ex:real line points}\n Let $\\mathcal{Q}$ be the additive $\\Bbbk$-linearization of $\\mathbb{R}$ as a category where paths move upwards.\n Let $r,s$ be positive real numbers and for each $x\\in r\\mathbb{Z}\\subset \\mathbb{R}$, let $\\mathcal P_x$ be the (unique) point relation in $\\mathcal{Q}$ through $x$ and $\\mathcal{I}$ the admissible ideal generated by $\\bigcup_{x\\in r\\mathbb{Z}} \\mathcal{P}_x$.\n Let $\\mathcal{J}$ be the the length relation in $\\mathcal{Q}\/\\mathcal{I}$ obtained by modding out by paths of length greater than $s$.\n By \\Cref{thm:length relations are admissible,thm:point relations are admissible} with \\Cref{lem:stack admissible} we obtain an admissible ideal $\\widetilde{\\Jay}$ given by the point relations at each $x\\in r\\mathbb{Z}$ and paths of length greater than $s$.\n \n If $r\\leq s$ then $\\mathcal{C}\/\\mathcal{I} = \\mathcal{C}\/\\widetilde{\\Jay}$ since we cannot have a morphism of length greater than $r$ in $\\mathcal{C}\/\\mathcal{I}$ anyway.\n If $r>s$ then we obtain paths of of length less than or equal to $s$ that do not pass through any $x\\in r\\mathbb{Z}$.\n The ``Auslander--Reiten space'' for the case $r>s$ is in \\Cref{fig:AR chopped mountains}; notice the similarity with \\Cref{fig:AR mountain range}.\n\\end{example}\n\n\\begin{example}[Complications with Cycles]\\label{ex: cycles length}\n For each $n\\in\\mathbb{N}$, let $\\mathcal{C}_n$ be circle whose radius is $\\frac{1}{2}e^{-n}$.\n Let $\\mathcal{C}$ be the additive $\\Bbbk$-linearization of $\\mathbb{R}_{\\leq 0}\\amalg(\\coprod_{n\\in\\mathbb{N}} \\mathcal{C}_n)\/ \\sim$, where $\\mathbb{R}\\ni -n\\sim 0\\in\\mathcal{C}_n$.\n See \\Cref{fig:cycles with decreasing length} for a visual depiction.\n \n We see that $\\mathcal{C}$ has length in $\\mathbb{R}_{\\geq 0}$.\n Let $\\mathcal{I}$ be a length relation.\n Since our length is in $\\mathbb{R}_{\\geq 0}$ we can say we are modding out by length $>L$ or $\\geq L$ for some $L>0\\in\\mathbb{R}$.\n \n Notice that for each $N\\in\\mathbb{N}$, there exists some $\\mathcal{C}_n$ with radius $r$ such that $Nr < L$.\n Therefore, there is no natural number $n$ such that for all nonisomorphism endomorphisms $f$ we have $f^n\\in\\mathcal{I}$.\n \n \\begin{figure}\n \\centering\n \\begin{tikzpicture}[scale=2, decoration={markings, mark=at position 0.6 with {\\arrow{>}}}]\n \\foreach \\x in {0, -1,...,-3}{\n \t\\node[outer sep=4+\\x] (\\x) at (\\x,0){$\\bullet$};\n \t\\node at (\\x.south){\\x};}\n \\node (-4) at (-4,0){$\\cdots$};\n \\foreach \\x [remember=\\x as \\y (initially 0)]in {-1, -2,...,-4}\n \t\\draw[very thick, postaction=decorate] (\\x.center)--(\\y.center);\n \\foreach \\x [evaluate=\\x as \\r using .5*e^(\\x\/2)] in {0, -1,...,-3}\n \\draw[radius=-\\r, very thick, postaction=decorate](\\x)+(0,\\r ) circle;\n \\end{tikzpicture}\n \\caption{Illustration of $\\mathcal{C}$ in \\Cref{ex: cycles length}, where we have glued smaller and smaller circles to each non-positive integer $n$ in $\\mathbb{R}_{\\leq 0}$.}\n \\label{fig:cycles with decreasing length}\n \\end{figure}\n\\end{example}\n\n\\begin{example}[Big wedge]\\label{ex:big wedge}\n Let $\\mathcal{C}$ be a cyclic continuous quiver of type $\\widetilde{A}$ as in \\cite{HR}.\n Let $\\widehat{\\mathcal{Q}} = (\\coprod_{\\mathbb{N}} \\mathcal{C} )\/\\sim$ where we join all the copies of the $\\mathcal{C}$ together at one point.\n Denote the wedge point by $X$.\n Let $\\mathcal{Q}$ be the additive $\\Bbbk$-linearization of $\\widehat{\\mathcal{Q}}$.\n Let us discuss the contruction of an admissible ideal out of point relations and a length relation.\n\n Notice $\\mathcal{Q}$ has length in $\\mathbb{R}_{\\geq 0}$.\n However, since the endomorphism ring of $X$ is an infinitely-generated monoid in $\\widehat{\\mathcal{Q}}$, we see that $\\End_{\\mathcal{C}}(X)\/ \\mathcal{I}(X,X)$ is not finite-dimensional.\n Instead, we must add a point relation on all but finitely-many different copies of $\\mathcal{C}$.\n If we do not have a point relation on all the cycles, we also need a length relation.\n Without such a combination, it is not possible to build an admissible ideal out of point relations and a length relation.\n\\end{example}\n\n\\subsection{The Real Plane}\\label{subsec:real plane}\n We now consider a continuous version of the grid quiver with commutativity relations as examined in \\cite{BBOS}.\n Let $\\widehat{\\mathcal{Q}}$ be the category whose objects are points in $\\mathbb{R}^2$.\n \n We now define the $\\Hom$ set between an arbitrary pair of points $(x,y)$ and $(z,w)$.\n Hom sets are given by considering paths made of up of horizontal and vertical line segments.\n For a pair $(x,y)$ and $(z,w)$, consider the set $P_{x,y}^{z,w}$ of all finite sequences $\\{(x_i,y_i)\\}_{i=i}^n$ such that\n \\begin{itemize}\n \\item $(x_1,y_1)=(x,y)$ and $(x_n,y_n)=(z,w)$,\n \\item $x_1\\leq x_2\\leq\\cdots \\leq x_n$ and $y_1\\leq y_2\\leq\\cdots \\leq y_n$,\n \\item $(x_i,y_i)\\neq (x_{i+1},y_{i+1})$ for all $1\\leq iz$ or $y>w$, then\n \\[ \\Hom_{\\widehat{\\mathcal{Q}}}((x,y), (z,w)) = \\emptyset.\\]\n If (i) $x=z$ and $y< w$ or (ii) $x< z$ and $y=w$, then\n \\[ \\Hom_{\\widehat{\\mathcal{Q}}}((x,y), (z,w)) = \\left\\{ \\{ (x,y), (z,w) \\}\\right\\}. \\]\n If $(x,y)=(z,w)$ then\n \\[ \\Hom_{\\widehat{\\mathcal{Q}}}((x,y), (x,y)) = \\left\\{ \\{ (x,y) \\}\\right\\}. \\]\n Composition is given by concatenating sequences and, if necessary, deleting a repeated term.\n \n Let $\\mathcal{Q}$ be the additive $\\Bbbk$-linearization of $\\widehat{\\mathcal{Q}}$ and let $(x,y),(z,w)\\in \\mathbb{R}^2$ such that $x\\varphi(n-1)$ and $k$. Hence, $\\sum(D_n+\\cdots+D_0)\\leq \\sum(C_{m+k}+\\cdots+C_0)$. Set $\\varphi(n)=m+k$. This proves the second inequality.\n\nFor the last inequality, since $C_{\\upharpoonright J}$ is left indecomposable, by Lemma \\ref{ncopies} we have $\\underline{2}.C_{\\upharpoonright J}\\leq C_{\\upharpoonright J}$ which implies that $\\sum \\underline{2}.C_{\\upharpoonright J}\\leq \\sum C_{\\upharpoonright J}$. \n\\end{proof} \n\n\nLemmas \\ref{Nonisoplc} and \\ref{ReducedequimorphP} imply the following. \n\n\\begin{theorem} \\label{Reduced}\nLet P be the sum of a countable labelled chain $C:=(I,\\ell)$ where $I$ has no least element and $\\ell(i)=(P_i,r(i))\\in \\mathcal{N}_{\\leq\\omega}\\times\\{-1, 0, +1\\}$ such that r takes 0 and $\\pm 1$ densely. Then $Sib(P)=2^{\\aleph_0}$. \n\\end{theorem}\n\n\n\n\n\n\\subsection{Direct Sum}\n \nThroughout this subsection $P$ is a countable direct sum of at least two non-empty connected $NE$-free posets. Thus, it is disconnected. \nIn this section we prove that $P$ has one or infinitely many siblings on condition that this property holds for each component of $P$. \n\n\\begin{lemma} \\label{Connecteddisconnected}\nIf some sibling of $P$ is connected, then $Sib(P)=\\infty$.\n\\end{lemma}\n\n\\begin{proof}\nLet $P'\\approx P$ where $P'$ is connected. So, $P'$ embeds into some component $Q$ of $P$. Since $P$ has at least two non-empty components, $P'\\oplus 1\\hookrightarrow Q\\oplus 1\\hookrightarrow P\\hookrightarrow P'$. So, for every $n$, $P'\\oplus \\Bar{K}_n\\hookrightarrow P'$ where $\\Bar{K}_n$ is an antichain of size $n$. Since $P'$ is connected and $P'\\approx P'\\oplus \\Bar{K}_n$, $P'$ and consequently $P$ has infinitely many pairwise non-isomorphic siblings. \n\\end{proof}\n\n\n\\begin{lemma} \\label{Infinitecomponent} \nIf some component of P has infinitely many siblings, then \\\\ $Sib(P)=\\infty$. \n\\end{lemma}\n\n\\begin{proof}\nLet $P$ satisfy the conditions of the lemma. Let $Q$ be a component of $P$ with infinitely many pairwise non-isomorphic siblings $Q_1, Q_2, Q_3, \\ldots$. For each $n$, let $P_n$ be the poset resulting from $P$ by replacing every component of $P$ which is equimorphic to $Q$ by $Q_n$. Now suppose that $P_m\\cong P_n$ for $m\\neq n$. Consider a component $Q'$ of $P_m$ equimorphic to $Q_m$. We know that $Q'$ is isomorphic to some component $Q''$ of $P_n$. We have $Q''\\cong Q'\\approx Q_m\\approx Q$ which implies that $Q''=Q_n$ because $Q''$ is a component of $P_n$. But then $Q_m\\cong Q_n$, a contradiction. \nTherefore, the resulting posets $P_n$ are pairwise non-isomorphic siblings of $P$. \n\\end{proof}\n\n\\begin{lemma} \\label{increasing}\nSuppose that P has infinitely many non-trivial components. Then $Sib(P)=\\infty$. \n\\end{lemma}\n\n\\begin{proof}\nSince $P$ has infinitely many non-trivial components and $\\mathcal{N}_{\\leq\\omega}$ is w.q.o, there is an increasing sequence $(P_n)_{n<\\omega}$ of non-trivial components of $P$. Let $\\mathcal{Q}$ be the direct sum of the non-trivial components of $P$ other than the $P_n$. We have $P=\\bigoplus_n P_n\\oplus \\mathcal{Q}\\oplus A$ where $A$ is the direct sum of the trivial components of $P$. Notice that since $P$ is countable, so is $A$. Therefore, $A$ embeds in $\\bigoplus_n P_{2n+1}$. Also $\\bigoplus_n P_n$ embeds into $\\bigoplus_n P_{2n}$. Hence, $P$ embeds into $P':=\\bigoplus_n P_n\\oplus \\mathcal{Q}$. That is, $P'\\approx P$. Similarly, $P'\\oplus \\Bar{K}_n \\approx P'\\approx P$ where $\\Bar{K}_n$ is an antichain of size $n$. Since $P'\\oplus\\Bar{K}_n$ has exactly $n$ trivial components, it follows that $P$ has infinitely many pairwise non-isomorphic siblings. \n\\end{proof}\n\nHence, by Lemmas \\ref{Connecteddisconnected}, \\ref{Infinitecomponent} and \\ref{increasing} we have the following. \n\n\\begin{proposition} \\label{InfinitesiblingnumberofP}\nLet P be a countable direct sum of NE-free posets with at least two non-empty components. If P is a sibling of some connected NE-free poset, or some component of P has infinitely many siblings, or P has infinitely many non-trivial components, then $Sib(P)=\\infty$. \n\\end{proposition}\n\n\n\n\n\\begin{lemma} \\label{Finitenontrivial} \nIf P has only finitely many non-trivial components and each component has only one sibling, then $Sib(P)=1$.\n\\end{lemma}\n\n\\begin{proof}\nSet $P:=\\bigoplus_{i0$ and that $P' \\subseteq P$ induces a sibling\nof $P$ via an embedding $f$. We prove that $f$ induces a bijection on the set of indices of components of $P$. First note that since the components of $P$ are connected, for each $i$, there is $j$ such that $f(P_i)\\subseteq P_j$. For each $i$, define $\\hat{f}(i)=j$ where $j$ is such that $f(P_i)\\subseteq P_j$. Suppose that for $i\\neq j$, $\\hat{f}(i)=\\hat{f}(j)=k$. It follows that $f$ embeds $P_i\\oplus P_j$ in $P_k$. We first show that $k$ cannot be $i$ or $j$. Suppose, without loss of generality, that $k=i$. Then $P_i\\oplus 1\\hookrightarrow P_i\\oplus P_j\\hookrightarrow P_i$ and by Lemma \\ref{Connecteddisconnected}, $P_i$ has infinitely many siblings, a contradiction. So, $k\\neq i, j$. It follows that $\\hat{f}(k)\\neq k$ because otherwise $f$ embeds $P_i\\oplus P_k$ in $P_k$ which implies that $P_k$ has infinitely many siblings, a contradiction. Further, $\\hat{f}(k)\\neq i, j$ because otherwise $P_i\\oplus P_j\\approx P_k$. By a similar argument, $\\hat{f}^2(k)\\neq i, j, k, \\hat{f}(k)$. Iterating, the $P_{\\hat{f}^n(k)}$ provide infinitely many non-trivial components of $P$, a contradiction. Thus, $\\hat{f}$ is injective on the set of indices of non-trivial components of $P$ and since there are finitely many such indices, $\\hat{f}$ is bijective. Extending $\\hat{f}$ to the set $I$ of indices of components of $P$, it follows that $\\hat{f}$ is bijective on $I$. Pick $i \\varphi(n-1)$ and some $k$. We know that the sequence $(a_n)_{n<\\omega}$ is coinitial in $J'$ and the $P_{a_n}$ are non-trivial. Pick a chain $a_{n_i}<\\cdots 0$ for every $i$. Now, suppose for two $i\\neq j$, $h(P(M_i)), h(P(M_j))\\subseteq P(M_k)$. By Lemma \\ref{Distinctobjects}, we have $\\lambda_i + \\lambda_j\\leq \\lambda_k$. This means that $\\lambda_k > \\lambda_i, \\lambda_j$. A contradiction is immediate when $k=i$ or $k=j$ because then $\\lambda_k > \\lambda_k$. Assume that $k\\neq i, j$. Since $\\lambda_k > \\lambda_i, \\lambda_j$, $P(M_k)$ cannot be embedded into $P(M_i)$ or $P(M_j)$. If $P(M_k)$ embeds in $P(M_k)$, then by Lemma \\ref{Distinctobjects}, $\\lambda_i+\\lambda_j+\\lambda_k \\leq \\lambda_k$, a contradiction to $\\lambda_i, \\lambda_j\\neq 0$. It follows that $P(M_k)$ embeds into some $P(M_l)$ where $l\\neq i, j, k$. Continuing this, we get infinitely many gs-connected children of $M$, a contradiction. Thus, $\\hat{h}$ is injective and since $X$ is finite, $\\hat{h}$ is a bijection on $X$. It also follows that $h(P_u)\\subseteq P_u$. Define $\\hat{h}(\\{u\\})=\\{u\\}$. \n\n\nSince $X$ is finite, for every $i$, there is some integer $m_i > 0$ such that $\\hat{h}^{m_i}(M_i)=M_i$. This means that for every $i$, $P(M_i)\\approx P(\\hat{h}.i(M_i))$ where $\\hat{h}.i$ is the orbit of $M_i$ under $\\hat{h}$. Hence, a sibling of $P(M)$ is of the form $\\bigoplus_iQ(M_i)\\oplus P'_u$ where $Q(M_i)\\approx P(M_i)$ for every $i$ and $P'_u\\cong P_u$ because $P_u$ has only one sibling. \n\\end{proof} \n\nFor case $v(M)=1$ where $M\\in T$ is edge minimal, we provide a result similar to Lemma \\ref{Bijectiononconnected}. \n\n\\begin{lemma} \\label{Bijectionondisconnected}\nLet $M\\in T$ be edge minimal such that $v(M)=1$ and $h$ an embedding of $P(M)$. \n\\begin{enumerate}\n \\item For every gs-disconnected child $M_i$ of $M$ we have $h(P(M_i))\\subseteq P(M_i)$.\n \\item For any chain $P_u^i$ in the representation $P_u^1 + P(M_1) + P_u^2 + P(M_2) + \\cdots + P(M_{k-1}) + P_u^k$ of $P(M)$ where $\\{u\\}$ is the unique possible gs-connected child of $M$, $h(P_u^i)\\subseteq P_u^i$. \n\\end{enumerate}\nHence, a sibling of $P(M)$ is of the form \n$$Q_u^1 + Q(M_1) + Q_u^2 + Q(M_2) + \\cdots + Q(M_{k-1}) + Q_u^k $$\nwhere $Q(M_i)\\approx P(M_i)$ and $Q_u^i\\approx P_u^i$ for every $i$. \n\\end{lemma}\n\n\\begin{proof}\nBy Proposition \\ref{Finiteps}, $P(M)$ is of the form \n$$P_u^1 + P(M_1) + P_u^2 + P(M_2) + \\cdots + P(M_{k-1}) + P_u^k\\ $$\nwhere each $M_i$ is a gs-disconnected child of $M$ and the $P_u^i$ are the intervals of the unique chain $P_u$ where $\\{u\\}$ is the possible gs-connected child of $M$. \n\n(1) Let $i$ and $x\\in P(M_i)$ be given. Since $M_i$ is gs-disconnected, take some $y$ in a component of $P(M_i)$ other than the component to which $x$ belongs. So, $x\\perp y$. We have $h(x)\\perp h(y)$. This implies that $h(x)\\notin P_u^j$ where $1\\leq j \\leq k$. Suppose that $h(x)\\in P(M_j)$ where $i\\neq j$. Without loss of generality assume that $P(M_i) <_P P(M_j)$. Since $h(x)\\perp h(y)$, we have $h(y)\\in P(M_j)$. So, for every $y\\in P(M_i)$ in a component other than the component to which $x$ belongs, we have $h(y)\\in P(M_j)$. Exchanging the role of $x$ and $y$, for every $x\\in P(M_i)$ in a component other than the component to which $y$ belongs, we have $h(x)\\in P(M_j)$. It means that $h(P(M_i))\\subseteq P(M_j)$. Let $\\lambda_i, \\lambda_j$ be the number of elements of $M_i, M_j$, respectively. By Lemma \\ref{Distinctobjects} we have $\\lambda_i \\leq \\lambda_j$. Thus, it is not the case that $P(M_j)$ embeds in $P(M_j)$ by $h$ because then we get $\\lambda_i+\\lambda_j \\leq \\lambda_j$, a contradiction to $\\lambda_i > 0$. So, $h(P(M_j))\\subseteq P(M_l)$ where $j < l$. Continuing this, we get infinitely many gs-disconnected children of $M$, a contradiction. \n\n(2) Let $i$ and some $x\\in P_u^i$ be given. Suppose $h(x)\\in P_u^j$ or $h(x)\\in P(M_j)$ where $j\\neq i$. Assume that $i < j$. The case $j 0$ there exists a dense and open subset $O^\\epsilon$ with the property that every $N \\in O^\\epsilon$ decomposes as $N \\cong A \\oplus B$ with $A$ indecomposable and $B$ $\\epsilon$-trivial.\n\\end{theorem}\n\nRecall that, in a topological space, a \\define{generic} property is one that holds for all points of some dense and open set.\nOur main result then implies that, for every $\\epsilon > 0$, the property of decomposing as a direct sum of an indecomposable and an $\\epsilon$-trivial module is generic.\n\nTo put this result into context, we mention a similar instance of a generic property of persistence modules. \nThe set of all finitely presentable one-parameter persistence modules that are \\emph{staggered} (meaning that the finite endpoints of the intervals in their barcode are all distinct) is dense but not open.\nNevertheless, for every $\\epsilon > 0$ there exists a dense and open set of modules that decomposes as $A \\oplus B$ with $A$ staggered and $B$ $\\epsilon$-trivial, as a direct consequence of the isometry theorem \\cite{lesnick,chazal-silva-glisse-oudot,bubenik-scott,bauer-lesnick}.\n\n\n\\smallskip\nWe prove \\cref{theorem:main-theorem} by combining the following two results, which are of independent interest.\nThe first one states that, in the space of finitely presentable two-parameter persistence modules, indecomposable modules are dense.\n\n\\begin{restatable}{proposition}{indecomposabledense}\n \\label{theorem:indecomposables-dense}\n Let $N : \\Rbf^2 \\to \\mathbf{vec}$ be finitely presentable.\n For every $\\epsilon > 0$, there exists an indecomposable persistence module $M : \\Rbf^2 \\to \\mathbf{vec}$ such that $d_I(M,N) \\leq \\epsilon$.\n\\end{restatable}\n\nThe second result relates the property of being close to an indecomposable to the property of decomposing as a direct sum of an indecomposable and a nearly trivial persistence module.\n\n\\begin{restatable}{proposition}{stabilityindecomposability}\n \\label{proposition:stability-indecomposability}\n Let $n \\geq 1$ and let $M : \\Rbf^n \\to \\mathbf{vec}$ be finitely presentable and indecomposable.\n For every $\\epsilon > 0$ there exists $\\delta > 0$ such that every persistence module $N : \\Rbf^n \\to \\mathbf{vec}$ with $d_I(M,N) < \\delta$ decomposes as $N \\cong A \\oplus B$ with $A$ indecomposable and $B$ $\\epsilon$-trivial.\n\\end{restatable}\n\n\\subparagraph{Discussion}\n\\cref{proposition:stability-indecomposability} motivates the following definition.\n\n\\begin{definition}\n Let $\\epsilon > 0$.\n A persistence module $M : \\Rbf^n \\to \\mathbf{vec}$ is \\define{$\\epsilon$-indecomposable} if we have $M \\cong A \\oplus B$ with $A$ indecomposable and $B$ $\\epsilon$-trivial.\n\\end{definition}\nOur main theorem then implies that, for every $\\epsilon > 0$, two-parameter persistence modules are generically $\\epsilon$-indecomposable.\nSince there are no general strong relationships between meager and measure-zero sets \\cite{oxtoby}, our results do not imply that random two-parameter persistence modules, according to some suitable probability measure, are $\\epsilon$-indecomposable.\nOne could then ask the following general questions:\nWhat are interesting and useful probability measures on spaces of multi-parameter persistence modules?\nAccording to these probability measures, how do random multi-parameter persistence modules decompose?\n\nWe prove \\cref{proposition:stability-indecomposability} in the generality of multi-parameter persistence modules, while we have given \\cref{theorem:indecomposables-dense} and consequently \\cref{theorem:main-theorem} only in the case of two-parameter persistence.\nWe believe the results hold for any number of parameters greater than one.\nSimilarly, the results extend to classes of modules more general than finitely presentable ones.\nWe leave these extensions for future work.\n\n\\subparagraph{Structure of the paper}\nIn \\cref{section:background}, we recall necessary background and introduce notation.\nIn \\cref{section:almost-indecomposable-open}, we prove \\cref{proposition:stability-indecomposability}.\nIn \\cref{section:indecomposables-dense}, we prove \\cref{theorem:indecomposables-dense} and \\cref{theorem:main-theorem}.\nIn \\cref{section:proof-main-lemma} we prove a key lemma that allows us to ``tack'' indecomposables together.\n\n\n\n\\ifarxivversion\n\\subparagraph{Acknowledgements}\nWe thank H\\r{a}vard Bjerkevik for interesting conversations and useful observations.\nL.~S.~thanks\nNicolas Berkouk,\nMathieu Carri\u00e8re,\nRen\u00e9 Corbet,\nChristian Hirsch,\nClaudia Landi,\nVadim Lebovici,\nDavid Loiseaux,\nEzra Miller,\nSteve Oudot,\nFran\u00e7ois Petit,\nand Alexander Rolle\nfor discussions during the \\emph{Metrics in Multiparameter Persistence workshop} (Lorentz Center, 2021).\nThis research has been conducted while U.~B.~was participating in the program \\emph{Representation Theory: Combinatorial Aspects and Applications};\nU.~B.~thanks the Centre for Advanced Study (CAS) at the Norwegian Academy of Science and Letters for their hospitality and support.\nThis research has been supported by the DFG Collaborative Research Center SFB\/TRR 109 \\emph{Discretization in Geometry and Dynamics}.\n\\fi\n\n\\section{Background}\n\\label{section:background}\n\nAlthough we use some notions from category theory, we only assume familiarity with basic concepts; in particular: categories, isomorphisms, functors, functor categories, direct sums, and kernels and cokernels.\nSee, e.g., \\cite{maclane} for an introduction.\n\nThroughout the paper, we fix a field $\\kbb$ and let $\\mathbf{vec}$ denote the category of finite dimensional $\\kbb$-vector spaces.\n\nAn \\define{extended metric space} consists of a set $A$ together with a function $d : A \\times A \\to \\Rbb \\cup \\{\\infty\\}$ such that, for all $a,b \\in A$ we have $d(a,b) = d(b,a) \\geq 0$ and $d(a,b) = 0$ if and only if $a=b$; and for all $a,b,c \\in A$ we have $d(a,c) \\leq d(a,b) + d(b,c)$, with the convention that $r + \\infty = \\infty + r = \\infty$ for all $r \\in \\Rbb \\cup \\{\\infty\\}$.\n\n\\subparagraph{Posets}\nWe let $\\Rbf$ denote the poset of real numbers with its standard ordering and reserve the notation $\\Rbb$ for the metric space of real numbers.\nLet $n \\in \\Nbb$ and consider the product poset $\\Rbf^n$.\nBy an abuse of notation, if $r \\in \\Rbf$, we interpret it as an element $r \\in \\Rbf^n$ all of whose coordinates are equal to $r$.\nThus, if for instance $s \\in \\Rbf^n$ and $r \\in \\Rbf$, then $s + r \\in \\Rbf^n$ denotes the element $(s_1 + r, s_2 + r, \\dots, s_n + r)$. \n\nWe interpret any poset $\\Pscr$ as a category with objects the elements of the poset and exactly one morphism $i \\to j \\in \\Pscr$ whenever $i \\leq j$.\n\n\\subparagraph{Persistence modules}\nLet $\\Pscr$ be a poset.\nA pointwise finite dimensional \\define{$\\Pscr$-persistence module} is a functor $M : \\Pscr \\to \\mathbf{vec}$.\nNote that all persistence modules in this paper are assumed to be pointwise finite dimensional and that, for the sake of brevity, we may omit this qualifier.\nWhen the indexing poset $\\Pscr$ is clear from the context, we may refer to a $\\Pscr$-persistence module simply as a persistence module or as a module.\nThe collection of all $\\Pscr$-persistence modules forms a category, where the morphisms are given by natural transformations.\n\nIf $M : \\Pscr \\to \\mathbf{vec}$ is a $\\Pscr$-persistence module and $i \\leq j \\in \\Pscr$, we let $\\phi^M_{i,j} : M(i) \\to M(j)$ denote the structure morphism corresponding to the morphism $i \\to j$ in $\\Pscr$ seen as a category.\n\nLet $\\Pscr$ be a poset and let $i \\in \\Pscr$.\nDefine the persistence module $\\Psf_i : \\Pscr \\to \\mathbf{vec}$ by\n\\[\n \\Psf_i(j) =\n \\begin{cases}\n \\kbb & i \\leq j\\\\\n 0 & \\text{else},\n \\end{cases}\n\\]\nwith all structure morphisms that are not forced to be zero being the identity $\\kbb \\to \\kbb$.\n\nA $\\Pscr$-persistence module is \\define{finitely presentable} if it is isomorphic to the cokernel of a morphism $\\bigoplus_{j \\in J} \\Psf_j \\to \\bigoplus_{i \\in I} \\Psf_i$, where $I$ and $J$ are finite multisets of elements of $\\Pscr$.\n\n\n\\subparagraph{Restrictions and extensions}\nIf $\\Qscr \\subseteq \\Pscr$ is an inclusion of posets and $M : \\Pscr \\to \\mathbf{vec}$ is a $\\Pscr$-persistence module, the \\define{restriction} of $M$ to $\\Qscr$, denoted $M|_\\Qscr : \\Qscr \\to \\mathbf{vec}$, is the $\\Qscr$-persistence module obtained by precomposing $M : \\Pscr \\to \\mathbf{vec}$ with the inclusion $\\Qscr \\to \\Pscr$.\n\nWe consider two main types of subposets of $\\Rbf^n$.\nIn one case, we let $\\{r_1 < r_2 < \\dots < r_k\\}$ be a finite set of real numbers and consider the product poset $\\{r_1 < r_2 < \\dots < r_k\\}^n \\subseteq \\Rbf^n$.\nWe refer to a subposet of $\\Rbf^n$ obtained in this way as a \\define{finite grid}.\nIn the other case, we let $\\{r_i\\}_{i \\in \\Zbb}$ be a countable set of real numbers without accumulation points and such that $r_i < r_{i+1}$ for all $i \\in \\Zbb$, and again consider the product poset $\\{r_i\\}^n \\subseteq \\Rbf^n$.\nWe refer to a subposet of $\\Rbf^n$ obtained in this way as a \\define{countable grid}.\nA \\define{regular grid} is any countable grid of the form $(s \\Zbf)^n = \\{s \\cdot m : m \\in \\Zbf\\}^n \\subseteq \\Rbf^n$ for $s > 0 \\in \\Rbb$, where $\\Zbf$ denotes the poset of integers.\nNote that, in the definitions of finite and countable grid, one could take a product of different subposets of $\\Rbf$, instead of an $n$-fold product of the same poset.\nSince we do not need this generality, we choose the more restrictive definition for simplicity.\n\n\nLet $\\Pscr \\subseteq \\Rbf^n$ be a finite or countable grid.\nGiven $M : \\Pscr \\to \\mathbf{vec}$ define $\\widehat{M} : \\Rbf^n \\to \\mathbf{vec}$\nby\n\\[\n \\widehat{M}(r) =\n \\begin{cases}\n M\\left(\\sup\\{ p \\in \\Pscr : p \\leq r\\}\\right) & \\text{if there exists $p \\in \\Pscr$ such that $p \\leq r$}\\\\\n 0 & \\text{else};\n \\end{cases}\n\\]\nfor its structure morphisms use the structure morphisms of $M$ and the fact that $\\sup\\{p \\in \\Pscr : p \\leq r\\} \\leq \\sup\\{p \\in \\Pscr : p \\leq s\\}$ whenever $r \\leq s$ and there exists $p \\in \\Pscr$ such that $p \\leq r$.\nWe refer to $\\widehat{M}$ as the \\define{extension} of $M$ along the inclusion $\\Pscr \\subseteq \\Rbf^n$.\nAs a side remark, we note that this notion of extension is an instance of the notion of left Kan extension from category theory, but we do not make use of this fact.\n\nIf $\\Pscr \\subseteq \\Rbf^n$ is a finite or countable grid and $M : \\Rbf^n \\to \\mathbf{vec}$, define the \\define{restriction-extension} of $M$ along $\\Pscr$, denoted $M_\\Pscr : \\Rbf^n \\to \\mathbf{vec}$, as the extension along $\\Pscr \\subseteq \\Rbf^n$ of the restriction $M|_\\Pscr : \\Pscr \\to \\mathbf{vec}$.\nGiven $M,N : \\Rbf^n \\to \\mathbf{vec}$ and a morphism $f : M \\to N$, there is a morphism $f_\\Pscr : M_\\Pscr \\to N_\\Pscr$ given by extending the restriction $f|_\\Pscr : M|_\\Pscr \\to N|_\\Pscr$.\nIt follows immediately that this construction is functorial in the sense that given modules $A,B,C : \\Rbf^n \\to \\mathbf{vec}$ and morphism $f : A \\to B$ and $g : B \\to C$, we have $g_\\Pscr \\circ f_\\Pscr = (g \\circ f)_\\Pscr : A_\\Pscr \\to C_{\\Pscr}$.\n\n\n\\begin{lemma}\n \\label{lemma:fp-is-extension-finite-poset}\n Let $n \\geq 1$ and let $M : \\Rbf^n \\to \\mathbf{vec}$.\n Then $M$ is finitely presentable if and only if there exists a finite grid $\\Pscr \\subseteq \\Rbf^n$ such that $M \\cong M_\\Pscr$.\n\\end{lemma}\n\\begin{proof}\n Assume that $M$ is finitely presentable, so that $M$ is isomorphic to the cokernel of a morphism $\\bigoplus_{j \\in J} \\Psf_j \\to \\bigoplus_{i \\in I} \\Psf_i$ with $I$ and $J$ finite subsets of $\\Rbf$.\n Consider the set $S = \\{x_k \\in \\Rbf : x \\in I \\cup J, \\, 1 \\leq k \\leq n\\}$ of all the coordinates of the points in $I \\cup J$, and the finite grid $\\Pscr = S^n \\subseteq \\Rbf^n$.\n It is straightforward to see that $M \\cong M_\\Pscr$.\n\n Now assume that $M \\cong \\widehat{N}$ with $\\Pscr \\subseteq \\Rbf^n$ a finite grid and $N : \\Pscr \\to \\mathbf{vec}$.\n Since the poset $\\Pscr$ has finitely many elements and $N$ is pointwise finite dimensional, there exists an epimorphism $e : \\bigoplus_{i \\in I} \\Psf_i \\to N$, for some finite multiset $I$ of elements of $\\Pscr$.\n Similarly, if $K$ is the kernel of the morphism $e$, there must exist an epimorphism $\\bigoplus_{j \\in J} \\Psf_j \\to K$, with $J$ finite.\n Putting these two morphisms together we see that $N$ is isomorphic to the cokernel of a morphism $\\bigoplus_{j \\in J} \\Psf_j \\to \\bigoplus_{i \\in I} \\Psf_i$.\n It is easy to see that $M$ is then isomorphic to the cokernel of the induced morphism $\\bigoplus_{j \\in J} \\widehat{\\Psf_j} \\to \\bigoplus_{i \\in I} \\widehat{\\Psf_i}$, and that $\\widehat{\\Psf_r} = \\Psf_r : \\Rbf^n \\to \\mathbf{vec}$ for all $r \\in \\Pscr$.\n\\end{proof}\n\nIt is straightforward to see that, in a grid extension persistence module, the structure maps that do not cross the grid are isomorphisms, as recorded in the following lemma.\n\n\\begin{lemma}\n \\label{lemma:structure-map-iso}\n Let $\\Pscr$ be a finite or countable grid and let $M : \\Rbf^n \\to \\mathbf{vec}$ be isomorphic to the extension of a $\\Pscr$-persistence module.\n Let $r < s \\in \\Rbf^n$ such that every $p \\in \\Pscr$ with $p \\leq s$ also satisfies $p \\leq r$.\n Then the structure morphism $\\phi^M_{r,s} : M(r) \\to M(s)$ is an isomorphism.\n \\qed\n\\end{lemma}\n\n\\subparagraph{Decomposition of persistence modules}\nThe proofs of the results in this section are standard, we include them in \\cref{appendix}, for completeness.\nLet $\\Pscr$ be a poset and let $M : \\Pscr \\to \\mathbf{vec}$.\nWe say that $M$ is \\define{decomposable} if there exist $A,B : \\Pscr \\to \\mathbf{vec}$ non-zero such that $M \\cong A \\oplus B$.\nIf $M$ is non-zero and not decomposable, we say that $M$ is \\define{indecomposable}.\n\n\nThe next two results follow from \\cite{botnan-crawleybovey} and the Krull--Remak--Schmidt--Azumaya theorem~\\cite{azumaya}.\n\n\n\\begin{restatable}{theorem}{theoremdecomposition}\n \\label{theorem:decomposition}\n Let $\\Pscr$ be a poset and let $M : \\Pscr \\to \\mathbf{vec}$ be a pointwise finite dimensional $\\Pscr$-persistence module.\n There exists a set $I$ and an indexed family of indecomposable $\\Pscr$-persistence modules $\\{M_i\\}_{i \\in I}$ such that $M \\cong \\bigoplus_{i \\in I} M_i$.\n Moreover, if $M \\cong \\bigoplus_{j \\in J} M_j$ for another indexed family of indecomposable $\\Pscr$-persistence modules $\\{M_j\\}_{j \\in J}$, then there exists a bijection $f : I \\to J$ such that $M_i \\cong M_{f(j)}$ for all $i \\in I$.\n\\end{restatable}\n\n\\begin{restatable}{lemma}{indecomposablelocalring}\n \\label{lemma:indecomposable-local-ring}\n Let $\\Pscr$ be a poset.\n A persistence module $M : \\Pscr \\to \\mathbf{vec}$ is indecomposable if and only if its endomorphism ring $\\End(M)$ is local.\n\\end{restatable}\n\n\\begin{proof}\n This is a direct consequence of \\cite[Theorem~1.1]{botnan-crawleybovey}.\n\\end{proof}\n\nThe following result states that direct sum decompositions behave well with respect to restrictions and extensions; its proof is straightforward.\n\n\\begin{lemma}\n \\label{lemma:decomposition-restriction-extension}\n Let $\\Pscr \\subseteq \\Rbf^n$ be a subposet.\n Let $M : \\Rbf^n \\to \\mathbf{vec}$ decompose as $M \\cong \\bigoplus_{i\\in I} M_i$ for some indexed family $\\{M_i\\}_{i \\in I}$ of persistence modules.\n Then $M|_\\Pscr \\cong \\bigoplus_{i\\in I} (M_i)|_\\Pscr$.\n Similarly, if $\\Pscr$ is a finite or countable grid and $M : \\Pscr \\to \\mathbf{vec}$ decomposes as $M \\cong \\bigoplus_{i\\in I} M_i$ for some indexed family $\\{M_i\\}_{i \\in I}$ of $\\Pscr$-persistence modules, then $\\widehat{M} \\cong \\bigoplus_{i\\in I} \\widehat{M_i}$.\n \\qed\n\\end{lemma}\n\nThe next result asserts that finitely presentable persistence modules decompose as a finite direct sum of finitely presentable modules.\nThis holds in the generality of persistence modules indexed by an arbitrary poset, but we do not need this generality.\n\n\\begin{restatable}{lemma}{decompositionfp}\n \\label{lemma:decomposition-fp}\n Let $n \\geq 1$ and let $M : \\Rbf^n \\to \\mathbf{vec}$.\n If $M$ is finitely presentable, then $M \\cong \\bigoplus_{i \\in I} M_i$ with $\\{M_i\\}_{i \\in I}$ a finite family of finitely presentable indecomposable $\\Rbf^n$-persistence modules.\n\\end{restatable}\n\n\\subparagraph{Interleavings, interleaving distance, and space of persistence modules}\nLet $n \\in \\Nbb$.\nIf $M : \\Rbf^n \\to \\mathbf{vec}$ is a $\\Rbf^n$-persistence module and $r \\in \\Rbf$, the \\define{$r$-shift} of $M$ is the persistence module $M[r] : \\Rbf^n \\to \\mathbf{vec}$ with $M[r](s) = M[s+r]$ for all $s \\in \\Rbf^n$ and $\\phi^{M[r]}_{s,t} = \\phi^M_{s+r,t+r}$ for all $s\\leq t \\in \\Rbf^n$.\nIf $r \\geq 0 \\in \\Rbf$, then there is a natural transformation $\\eta^M_r : M \\to M[r]$ with component $M(a) \\to M[r](a) = M(a+r)$ given by the structure morphism $\\phi^M_{a,a+r}$.\n\nLet $M,N : \\Rbf^n \\to \\mathbf{vec}$ and $\\epsilon \\geq 0 \\in \\Rbf$.\nAn \\define{$\\epsilon$-interleaving} between $M$ and $N$ consists of a pair of morphisms $f : M \\to N[\\epsilon]$ and $g : N \\to M[\\epsilon]$ such that $g[\\epsilon] \\circ f = \\eta^M_{2\\epsilon}$ and $f[\\epsilon] \\circ f = \\eta^N_{2\\epsilon}$.\nThe \\define{interleaving distance} between $M$ and $N$ is\n\\[\n d_I(M,N) = \\inf\\left(\\{\\epsilon \\geq 0 : \\text{ exists an $\\epsilon$-interleaving between $M$ and $N$ }\\} \\cup \\{\\infty\\}\\right) \\in \\Rbb \\cup \\{\\infty\\}.\n\\]\n\nBy composing interleavings, one shows that $d_I$ satisfies the triangle inequality.\nUsing \\cref{lemma:structure-map-iso}, one sees that if $M, N : \\Rbf^n \\to \\mathbf{vec}$ are finitely presentable and $d_I(M,N) = 0$, then $M$ and $N$ are isomorphic \\cite[Corollary~6.2]{lesnick}.\nThis implies that $d_I$ defines an extended metric on the set of isomorphism classes of finitely presentable $\\Rbf^n$-persistence modules.\nWe mention here that the collection of all isomorphism classes of finitely presentable $\\Rbf^n$-persistence modules is indeed a set, as opposed to a proper class, but we shall not delve into the details as they are standard and not relevant to the results presented here.\n\nWe consider the set of isomorphism classes of finitely presentable $\\Rbf^n$-persistence modules which we topologize with the topology generated by the open balls with respect to the interleaving distance.\n\nLet $\\epsilon > 0$.\nA persistence module $M : \\Rbf^n \\to \\mathbf{vec}$ is \\define{$\\epsilon$-trivial} if, for every $r \\in \\Rbf^n$, the structure morphism $\\phi^M_{r,r+\\epsilon} : M(r) \\to M(r + \\epsilon)$ is the zero morphism.\nNote that this is equivalent to $M$ being $\\epsilon\/2$-interleaved with the zero module.\nThus, if $d_I(M,0) < \\epsilon\/2$, then $M$ is $\\epsilon$-trivial.\n\n\n\\subparagraph{Shifts and restriction-extensions}\nIf $\\Pscr = (\\{s_i\\}_{i \\in \\Zbb})^n \\subseteq \\Rbf^n$ is a countable grid and $r \\in \\Rbf$, define the \\define{shifted} countable grid $\\Pscr + r = (\\{s_i + r\\}_{i \\in \\Zbb})^n \\subseteq \\Rbf^n$.\nLet $M : \\Rbf^n \\to \\mathbf{vec}$ and $s \\in \\Rbf$.\nIf $a \\in \\Rbf^n$, then $\\sup\\{ p + r : p + r \\leq a + s\\} = \\sup\\{p + r - s : p + r - s \\leq a\\} + s$.\nThus,\n\\[\n M_{\\Pscr + r}[s] = M[s]_{\\Pscr + (r-s)},\n\\]\na fact that we use repeatedly in \\cref{section:almost-indecomposable-open}.\nIn particular, given $N : \\Rbf^n \\to \\mathbf{vec}$ and a morphism $f : M \\to N[s]$, we get a morphism\n\\begin{equation}\n \\label{equation:shift-and-restriction-extension}\n f_\\Pscr : M_{\\Pscr + r} \\to N[s]_{\\Pscr + r} = N_{\\Pscr + r + s}[s].\n\\end{equation}\n\n\\section{Nearly indecomposables are open}\n\\label{section:almost-indecomposable-open}\n\nBefore we can prove \\cref{proposition:stability-indecomposability}, we need a definition and a lemma.\n\n\\begin{definition}\n Let $\\alpha > 0$.\n A countable grid $\\Pscr = \\left(\\{s_i\\}_{i \\in \\Zbb}\\right)^n \\subseteq \\Rbf^n$ has \\define{mesh width $\\alpha$} if for all $i \\in \\Zbb$ we have $s_{i+1} - s_i \\leq \\alpha$.\n\\end{definition}\n\nNote that, for any fixed $\\alpha > 0$, every countable grid $S^n = \\left(\\{s_i\\}_{i \\in \\Zbb}\\right)^n$ is included in some countable grid with mesh width $\\alpha$, for example, the countable grid $(S \\cup (\\alpha\\Zbf))^n$.\n\n\\begin{lemma}\n \\label{lemma:factor-bounded-grid}\n Let $\\alpha > 0$, let $\\Pscr$ be a countable grid with mesh width $\\alpha$, and let $L : \\Rbf^n \\to \\mathbf{vec}$.\n For every $r < \\alpha$ there exists a morphism $m : L_{\\Pscr + r}[r] \\to L_\\Pscr[\\alpha]$ rendering the following square commutative:\n \\[\n \\begin{tikzpicture}\n \\matrix (m) [matrix of math nodes,row sep=4em,column sep=7em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { L_{\\Pscr} & L_{\\Pscr}[\\alpha] \\\\\n L[r]_\\Pscr & L_{\\Pscr + r}[r],\\\\};\n \\path[line width=0.75pt, -{>[width=8pt]}]\n (m-1-1) edge node [above] {$\\eta^{L_\\Pscr}_\\alpha$} (m-1-2)\n (m-1-1) edge node [left] {$(\\eta^L_r)_\\Pscr$} (m-2-1)\n (m-2-2) edge node [right] {$m$} (m-1-2)\n (m-2-1) edge [-,double equal sign distance] (m-2-2)\n ;\n \\end{tikzpicture}\n \\]\n\\end{lemma}\n\\begin{proof}\n We start by defining the morphism $m$, which is equivalently a morphism $m : L_{\\Pscr+r} \\to L_\\Pscr[\\alpha-r]$.\n If $a \\in \\Rbf^n$, then\n $L_{\\Pscr + r}(a) = L(\\sup\\{ s + r : s \\in \\Pscr, s + r \\leq a\\})$ and\n \\[\n L_{\\Pscr}[\\alpha-r](a) = L_{\\Pscr}(a + \\alpha - r) = L(\\sup\\{s : s \\in \\Pscr, s \\leq a + \\alpha - r\\}).\n \\]\n Let $s_0$ such that $s_0 + r = \\sup\\{ s + r : s \\in \\Pscr, s + r \\leq a\\}$ so that $L_{\\Pscr + r}(a) = L(s_0 + r)$.\n If $\\Pscr = (\\{s_i\\}_{i \\in \\Zbb})^n$, then $s_0 = (s_{i_1}, \\dots, s_{i_n})$.\n Let $s_1 = (s_{i_1+1}, \\dots, s_{i_n+1})$.\n Note that, since $\\Pscr$ has mesh width $\\alpha$, we have $s_1 - \\alpha \\leq s_0$.\n Thus, $s_1 - \\alpha + r \\leq a$, which implies $s_1 \\leq a + \\alpha - r$.\n This means that $s_1 \\leq \\sup\\{s : s \\in \\Pscr, s \\leq a + \\alpha - r\\}$.\n We can then consider the composite:\n \\[\n L_{\\Pscr + r}(a) = L(s_0) \\to L(s_1) \\to L(\\sup\\{s : s \\in \\Pscr, s \\leq a + \\alpha - r\\}) = L_{\\Pscr}[\\alpha-r](a),\n \\]\n given by the structure morphisms of $L$.\n Let us denote the above composite by $m_a$.\n Since the morphism was constructed only using the structure morphisms of $L$, it is straightforward to check that the morphisms $m_a : L_{\\Pscr + r}(a) \\to L_{\\Pscr}[\\alpha-r](a)$ assemble into a natural transformation $m : L_{\\Pscr+r} \\to L_\\Pscr[\\alpha-r]$.\n By the same reason, it follows that $m$ renders commutative the square in the statement.\n\\end{proof}\n\n\n\\stabilityindecomposability*\n\\begin{proof}[Proof outline]\\renewcommand{\\qedsymbol}{}\n Let $M$ be as in the statement.\n Let us start by giving an outline of the proof strategy.\n We start by arguing that the persistence module $M$ is isomorphic to the extension of a persistence module over a sufficiently fine countable grid.\n We then choose a sufficiently small $\\delta > 0$ and consider an arbitrary module $N$ at interleaving distance at most $\\delta$ from $M$.\n Next, we use \\cref{lemma:structure-map-iso}, which says that the extension of a module over a grid can only change when one of its coordinates crosses a point in the grid; this is used to show that, after restricting to appropriate grids, a restriction of $M$ is indecomposable and isomorphic to a direct summand of a restriction of $N$.\n This allows us to find a decomposition of $N$ into two summands, one of which is indecomposable and is related to a restriction of $M$.\n Finally, we show that the other summand is close to the zero module in the interleaving distance.\n\\end{proof}\n\n\n\\begin{proof}\n By \\cref{lemma:fp-is-extension-finite-poset}, the module $M$ is isomorphic to the extension of its restriction to some finite grid $\\Qscr = \\{r_1 < r_2 < \\dots < r_k\\}^n \\subseteq \\Rbf^n$.\n Let\n \\[\n \\tau = \\min_{1\\leq i \\leq k-1} r_{i+1} - r_i\\;\\;, \\;\\; \\alpha = \\min\\left(\\epsilon\/4\\; , \\tau\\right)\\;\\;, \\;\\; \\delta < \\alpha\/4,\n \\]\n and let $\\Pscr = (\\{s_i\\}_{i \\in \\Zbb})^n$ be a countable grid with mesh width $\\alpha$ and containing $\\Qscr$.\n Note that $M$ is isomorphic to the extension of its restriction to $\\Pscr$, that is $M \\cong M_\\Pscr$.\n\n Assume given $N : \\Rbf^n \\to \\mathbf{vec}$ and a $\\delta$-interleaving $f : M \\to N[\\delta]$ and $g : N \\to M[\\delta]$ between $M$ and $N$.\n The interleaving equalities $g[\\delta] \\circ f = \\eta^M_{2\\delta}$ and $f[2\\delta] \\circ g[\\delta] = \\eta^{N[\\delta]}_{2\\delta}$ induce, by \\cref{equation:shift-and-restriction-extension}, the following commutative diagram of $\\Rbf^n$-persistence modules:\n \\begin{equation}\n \\label{equation:decomposition-N}\n \\begin{tikzpicture}[baseline=(current bounding box.center)]\n \\matrix (m) [matrix of math nodes,row sep=4em,column sep=3em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { & N_{\\Pscr + \\delta}[\\delta] & & N_{\\Pscr + 3\\delta}[3\\delta] \\\\\n M_{\\Pscr} & & M_{\\Pscr + 2\\delta}[2\\delta] & \\\\};\n \\path[line width=0.75pt, -{>[width=8pt]}]\n (m-2-1) edge node [above] {$\\cong$} node [below] {$(\\eta^M_{2\\delta})_{\\Pscr}$} (m-2-3)\n (m-2-1) edge [>->] node [above left] {$f_{\\Pscr}$} (m-1-2)\n (m-1-2) edge [->>] node [above right] {$g[\\delta]_{\\Pscr}$} (m-2-3)\n (m-1-2) edge node [above] {$(\\eta^{N[\\delta]}_{2\\delta})_{\\Pscr}$} (m-1-4)\n (m-2-3) edge node [below right] {$f[2\\delta]_\\Pscr$} (m-1-4)\n ;\n \\end{tikzpicture}\n \\end{equation}\n The bottom horizontal morphism is an isomorphism by \\cref{lemma:structure-map-iso} and the fact that $2\\delta < \\alpha\/2 \\leq \\tau\/2$.\n This implies that the left diagonal morphism is a split monomorphism and that the middle diagonal morphism is a split epimorphism.\n\n In particular, it follows that $M_{\\Pscr}$ is a direct summand of $N_{\\Pscr+\\delta}[\\delta]$, so that $N_{\\Pscr+\\delta}[\\delta] \\cong M_{\\Pscr} \\oplus X$ for some $X$.\n At the same time, by \\cref{theorem:decomposition}, there exists a decomposition $N \\cong \\bigoplus_{i \\in I} N_i$ with $N_i$ indecomposable for all $i \\in I$.\n Using \\cref{lemma:decomposition-restriction-extension}, we see that, by restricting to $\\Pscr$ and extending, we get a decomposition $N_{\\Pscr+\\delta} \\cong \\bigoplus_{i \\in I} (N_i)_{\\Pscr+\\delta}$, where now the summands $\\{(N_i)_{\\Pscr+\\delta}\\}_{i \\in I}$ may not be indecomposable anymore.\n Nevertheless, since $M_\\Pscr \\cong M$ is indecomposable by assumption, there exists $i \\in I$ such that $(N_i)_{\\Pscr+\\delta}\\cong M_\\Pscr[-\\delta] \\oplus X'$ for some $X'$, using the fact that $N_{\\Pscr+\\delta}[\\delta] \\cong M_{\\Pscr} \\oplus X$ and \\cref{theorem:decomposition}.\n We consider the decomposition\n \\[\n N_{\\Pscr + \\delta}\\;\\; \\cong \\;\\;(N_i)_{\\Pscr + \\delta} \\oplus \\bigoplus_{j \\in I \\setminus i} (N_j)_{\\Pscr + \\delta} \\;\\;\\cong \\;\\;(M_\\Pscr[-\\delta] \\oplus X') \\oplus \\bigoplus_{j \\in I \\setminus i} (N_j)_{\\Pscr + \\delta},\n \\]\n and let $A = N_i$ and $B = \\bigoplus_{j \\in I \\setminus i} N_j$.\n Since this will be of use later, we note here that, by construction, $B_\\Pscr$ is isomorphic to a summand of $X$.\n \n We have thus decomposed $N \\cong A \\oplus B$ with $A$ indecomposable.\n To conclude, it remains to show that $B$ is $\\epsilon$-trivial, and for this it is enough to show that $d_I(B,0) < \\epsilon\/2$.\n By definition of $\\Pscr$, we have $d_I(B,B_\\Pscr) < \\alpha \\leq \\epsilon\/4$, which reduces the problem to showing that $d_I(B_\\Pscr,0) < \\epsilon\/4$, by the triangle inequality.\n Since $B_\\Pscr$ is a summand of $X$, it is enough to prove that $d_I(X,0) < \\epsilon\/4$.\n We will show that $\\eta^X_{\\alpha} : X \\to X[\\alpha]$ is the zero morphism, which implies that $d_I(X,0) \\leq \\alpha\/2 \\leq \\epsilon\/8 < \\epsilon\/4$.\n\n By the naturality of shifting, that is, by the commutativity of the following square\n \\[\n \\begin{tikzpicture}\n \\matrix (m) [matrix of math nodes,row sep=4em,column sep=7em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { M_\\Pscr \\oplus X & M_\\Pscr[\\alpha] \\oplus X[\\alpha] \\\\\n N_{\\Pscr + \\delta} [\\delta] & N_{\\Pscr + \\delta}[\\delta + \\alpha],\\\\};\n \\path[line width=0.75pt, -{>[width=8pt]}]\n (m-1-1) edge node [above] {$\\left(\\eta^{M_\\Pscr}_\\alpha, \\eta^X_\\alpha\\right)$} (m-1-2)\n (m-1-1) edge node [left] {$\\cong$} (m-2-1)\n (m-1-2) edge node [left] {$\\cong$} (m-2-2)\n (m-2-1) edge node [above] {$\\eta^{N_{\\Pscr+\\delta}[\\delta]}_\\alpha$} (m-2-2)\n ;\n \\end{tikzpicture}\n \\]\n it is sufficient to show that the following composite is the zero morphism:\n \\begin{equation}\n \\label{equation:sufficient-zero-morphism}\n X \\rightarrowtail M_{\\Pscr} \\oplus X \\xrightarrow{\\cong} N_{\\Pscr+\\delta}[\\delta] \\xrightarrow{\\eta^{N_{\\Pscr+\\delta}[\\delta]}_\\alpha} N_{\\Pscr+\\delta}[\\delta + \\alpha].\n \\end{equation}\n Note that, by the construction of the splitting $N_{\\Pscr+\\delta}[\\delta] \\cong M_{\\Pscr} \\oplus X$ and the commutativity of the right triangle in \\cref{equation:decomposition-N}, the following composite is the zero morphism\n \\begin{equation}\n \\label{equation:zero-morphism}\n X \\rightarrowtail M_{\\Pscr} \\oplus X \\xrightarrow{\\cong} N_{\\Pscr+\\delta}[\\delta]\n \\xrightarrow{(\\eta^{N[\\delta]}_{2\\delta})_{\\Pscr}} N_{\\Pscr + 3\\delta}[3\\delta].\n \\end{equation}\n Thus, it suffices to see that the right-most morphism in \\cref{equation:sufficient-zero-morphism} factors through the right-most morphism in \\cref{equation:zero-morphism}.\n Letting $L = N[\\delta]$ and using \\cref{equation:shift-and-restriction-extension},\n this translates to the claim that the morphism \n $\\eta^{L_{\\Pscr}}_\\alpha : L_{\\Pscr} \\to L_{\\Pscr+\\delta}[\\delta + \\alpha]$ factors through $(\\eta^{L}_{2\\delta})_{\\Pscr} : L_{\\Pscr} \\to L_{\\Pscr + 2\\delta}[2\\delta]$, which follows from \\cref{lemma:factor-bounded-grid} by letting $r = 2\\delta = \\alpha\/2 < \\alpha$.\n\\end{proof}\n\n\\begin{corollary}\n \\label{corollary:open-with-indecomposables}\n Let $n \\geq 1$.\n For every $\\epsilon \\geq 0$, there exists an open set $O^\\epsilon$ of isomorphism classes of finitely presentable $n$-parameter persistence modules, containing the set of all indecomposable finitely presentable modules as a subset, and such that every $N \\in O^\\epsilon$ decomposes as $N \\cong A \\oplus B$ with $A$ indecomposable and $d_I(B,0) < \\epsilon$.\n\\end{corollary}\n\\begin{proof}\n Fix $\\epsilon > 0$.\n By \\cref{proposition:stability-indecomposability}, for every indecomposable finitely presentable $M : \\Rbf^n \\to \\mathbf{vec}$ there exists $\\delta^M > 0$ such that, for all finitely presentable $N : \\Rbf^n \\to \\mathbf{vec}$ in the open $\\delta^M$-ball around $M$ in the interleaving distance, it holds that $N$ decomposes as $N \\cong A \\oplus B$ with $A$ indecomposable and $B$ $\\epsilon$-trivial.\n It is then clear that the set\n \\[\n O^\\epsilon := \\bigcup_{\\substack{M \\text{ finitely presentable} \\\\ \\text{and indecomposable}}} \\{N : \\text{ $N$ is finitely presentable and } d_I(M,N) < \\delta^M \\}\n \\]\n satisfies the conditions in the statement.\n\\end{proof}\n\n\n\n\\section{Indecomposables are dense}\n\\label{section:indecomposables-dense}\n\n\nThe proof of \\cref{theorem:indecomposables-dense} depends on the following key lemma, which lets us ``tack together'' two indecomposable modules to obtain a third indecomposable module that is at small interleaving distance from the direct sum of the initial modules.\nThe proof is in \\cref{section:proof-main-lemma}.\n\n\\begin{restatable}[Tacking Lemma]{lemma}{mainlemma}\n \\label{lemma:main-lemma-attaching}\n Let $A,B : \\Rbf^2 \\to \\mathbf{vec}$ be finitely presentable, indecomposable, and isomorphic to extensions of $\\Pscr$-persistence modules for some regular grid $\\Pscr \\subseteq \\Rbf^2$.\n For every $\\delta > 0$ there exists a regular grid $\\Qscr \\supseteq \\Pscr$ and $M : \\Rbf^2 \\to \\mathbf{vec}$ with $M$ indecomposable, finitely presentable, isomorphic to the extension of a $\\Qscr$-persistence module, and such that\n \\[\n \\mathsf{im}\\left(\\eta^M_\\delta\\right) \\cong \\mathsf{im}\\left(\\eta^A_\\delta\\right) \\oplus \\mathsf{im}\\left(\\eta^B_\\delta\\right).\n \\]\n\\end{restatable}\n\nWe also need the following two simple lemmas.\n\n\\begin{lemma}\n \\label{lemma:image-to-interleaving}\n Let $A : \\Rbf^n \\to \\mathbf{vec}$ and let $\\delta \\geq 0$.\n Consider $\\mathsf{im}(\\eta^A_\\delta ) : \\Rbf^n \\to \\mathbf{vec}$, the image of the morphism of persistence modules $\\eta^A_\\delta : A \\to A[\\delta]$ given by the structure morphisms.\n Then $A$ and $\\mathsf{im}(\\eta^A_\\delta)$ are $\\delta$-interleaved.\n As a consequence, if $B : \\Rbf^n \\to \\mathbf{vec}$ satisfies $\\mathsf{im}(\\eta^A_\\delta) \\cong \\mathsf{im}(\\eta^B_\\delta)$, then $A$ and $B$ are $2\\delta$-interleaved.\n\\end{lemma}\nIn fact, the $2\\delta$ can be improved to $\\delta$ using the notion of asymmetric interleaving \\cite[Section~2.6.1]{lesnick-thesis}, but this is not necessary for our purposes.\n\\begin{proof}\n The second statement follows directly from the first one.\n By construction, there is a canonical epimorphism $A \\to \\mathsf{im}(\\eta^A_\\delta)$, and thus we get a morphism $f : A \\to \\mathsf{im}(\\eta^A_\\delta)[\\delta]$, by composing with $\\eta^{\\mathsf{im}(\\eta^A_\\delta)}_\\delta: \\mathsf{im}(\\eta^A_\\delta) \\to \\mathsf{im}(\\eta^A_\\delta)[\\delta]$.\n Similarly, there is a canonical monomorphism $g : \\mathsf{im}(\\eta^A_\\delta) \\to A[\\delta]$.\n A straightforward check shows that $f$ and $g$ form a $\\delta$-interleaving.\n\\end{proof}\n\n\\begin{lemma}\n \\label{lemma:eta-direct-sum}\n Let $A,B : \\Rbf^n \\to \\mathbf{vec}$ and $\\epsilon \\geq 0$.\n Then $\\mathsf{im}\\left(\\eta^A_\\epsilon\\right) \\oplus\\mathsf{im}\\left(\\eta^B_\\epsilon\\right) \\cong \\mathsf{im}\\left(\\eta^{A \\oplus B}_\\epsilon\\right)$.\n \\qed\n\\end{lemma}\n\n\\indecomposabledense*\n\\begin{proof}\n We start by reducing the problem to the case in which the given module is an extension of a module defined over a regular grid.\n Consider the regular grid $\\Pscr = ((\\epsilon\/2)\\Zbf)^2$ and note that $d_I(N_\\Pscr,N) \\leq \\epsilon\/2$.\n Let $L = N_\\Pscr$.\n It is easy to see that $L$ is finitely presentable.\n\n If $L = 0$, then take $M$ to be the persistence module such that $M(x,y) = \\kbb$ if $0 \\leq x < \\epsilon$ and $y \\geq 0$, and $0$ otherwise, with all the structure morphisms that are not forced to be the zero morphism being the identity $\\kbb \\to \\kbb$.\n It is straightforward to see that $M$ is indecomposable and that $d_I(M,0) = \\epsilon\/2$.\n This case now follows from $d_I(N,M) \\leq d_I(N,L) + d_I(L,M) \\leq \\epsilon\/2 + d_I(0,M) \\leq \\epsilon$.\n\n We now assume that $L$ is non-zero.\n By \\cref{lemma:decomposition-fp}, there exist finitely presentable indecomposables $X_1, \\dots, X_k : \\Rbf^2 \\to \\mathbf{vec}$ such that $L \\cong X_1 \\oplus \\dots \\oplus X_k$.\n Note that, for all $1 \\leq i \\leq k$, the persistence module $X_i$ is isomorphic to the extension of a $\\Qscr$-persistence module for any grid $\\Qscr \\supseteq \\Pscr$.\n Let us define $M_i : \\Rbb^2 \\to \\mathbf{vec}$ for $1 \\leq i \\leq k$ inductively.\n Let $M_1 = X_1$ and, for $2 \\leq i \\leq k$, let $M_{i}$ be obtained by setting $A = X_i$, $B = M_{i-1}$, and $\\delta = \\epsilon\/4$ in \\cref{lemma:main-lemma-attaching}.\n Thus, $M_i$ is finitely presentable, indecomposable, and satisfies\n \\[\n \\mathsf{im}\\left(\\eta^{M_i}_{\\epsilon\/4}\\right) \\cong \\mathsf{im}\\left(\\eta^{X_i}_{\\epsilon\/4}\\right) \\oplus \\mathsf{im}\\left(\\eta^{M_{i-1}}_{\\epsilon\/4}\\right).\n \\]\n Let $M = M_k$.\n It follows by induction and \\cref{lemma:eta-direct-sum} that \n \\[\n \\mathsf{im}\\left(\\eta^{M}_{\\epsilon\/4}\\right) \\cong \\mathsf{im}\\left(\\eta^{X_1}_{\\epsilon\/4}\\right) \\oplus \\dots \\oplus \\mathsf{im}\\left(\\eta^{X_k}_{\\epsilon\/4}\\right) \\cong \\mathsf{im}\\left(\\eta^{X_1 \\oplus \\dots \\oplus X_k}_{\\epsilon\/4}\\right) \\cong \\mathsf{im}\\left(\\eta^L_{\\epsilon\/4}\\right).\n \\]\n Using \\cref{lemma:image-to-interleaving}, we get that $M$ and $L$ are $\\epsilon\/2$-interleaved.\n By the triangle inequality, we get $d_I(M,N) \\leq d_I(M,L) + d_I(L,N) \\leq \\epsilon$, as required.\n\\end{proof}\n\nWe can now give the proof of our main result.\n\n\\begin{proof}[Proof of \\cref{theorem:main-theorem}]\n Given $\\epsilon > 0$, the set $O^\\epsilon$ of \\cref{corollary:open-with-indecomposables} is open and contains all indecomposables finitely presentable modules.\n Since, when $n=2$, the set of indecomposables is dense by \\cref{theorem:indecomposables-dense}, the result follows.\n\\end{proof}\n\n\n\\section{Tacking indecomposables together}\n\\label{section:proof-main-lemma}\n\nIn this section, we prove \\cref{lemma:main-lemma-attaching} (the Tacking Lemma).\nWe first summarize the strategy.\n\n\n\\begin{proof}[Proof outline]\\renewcommand{\\qedsymbol}{}\nWe start by defining a persistence module $\\Gsf$ on a finite grid which allows us to ``tack together'' indecomposable modules.\nNow, fix modules $A$ and $B$ as in the statement.\nThe tacking works in three steps.\nIn \\cref{lemma:add-1-dim-corner}, we replace $A$ and $B$ by modules which admit a ``thin corner'' (\\cref{definition:1-dim-corner}); the purpose of this thin corner is that it allows us to attach a copy of $\\Gsf$ to each of the modules, which we do in the next step.\nIn \\cref{lemma:add-antenna-attachment}, we replace the modules by modules that admit a ``horizontal antenna attachment'' (\\cref{definition:horizontal-antenna-attachment}); the purpose of this horizontal antenna attachment is that it allows us to tack together the two modules with a third copy of $\\Gsf$. \nFinally, in \\cref{lemma:actual-tacking}, we construct $M$ by tacking together the two modules.\n\\cref{figure:diagram-proof-main-lemma} shows a diagrammatic description of these steps.\n\\end{proof}\n\n\\begin{figure}[h]\n \\centering\n \\includegraphics[width=1\\linewidth]{pictures\/diagram-steps.eps}\n \\caption{A schematic summary of the main steps in the proof of the tacking lemma.\n The transition from (0.) to (1.) corresponds to \\cref{lemma:add-1-dim-corner}; the transition from (1.) to (2.) corresponds to \\cref{lemma:add-antenna-attachment}; the transition from (2.) to (3.) corresponds to \\cref{lemma:actual-tacking}.}\n \\label{figure:diagram-proof-main-lemma}\n\\end{figure}\n\n\\begin{definition}\n\\label{def:G}\n Let $\\Pscr = \\{0, 1, 2, 3, 4\\}^2$ and define $\\Gsf : \\Pscr \\to \\mathbf{vec}$ as follows:\n \\[ \\footnotesize\n \\begin{tikzpicture}\n \\matrix (m) [matrix of math nodes,row sep=3em,column sep=3em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { \\kbb & \\kbb & \\kbb & \\kbb & \\kbb \\\\\n \\kbb & \\kbb^2 & \\kbb^2 & \\kbb^2 & \\kbb \\\\\n 0 & \\kbb & \\kbb^2 & \\kbb^2 & \\kbb \\\\\n 0 & 0 & \\kbb & \\kbb^2 & \\kbb \\\\\n 0 & 0 & 0 & \\kbb & \\kbb \\\\};\n \\path[line width=0.5pt, -{>[width=6pt]}]\n (m-1-1) edge [-,double equal sign distance] (m-1-2)\n (m-1-2) edge [-,double equal sign distance] (m-1-3)\n (m-1-3) edge [-,double equal sign distance] (m-1-4)\n (m-1-4) edge [-,double equal sign distance] (m-1-5)\n\n (m-2-2) edge [-,double equal sign distance] (m-2-3)\n (m-2-3) edge [-,double equal sign distance] (m-2-4)\n\n (m-3-3) edge [-,double equal sign distance] (m-3-4)\n\n (m-5-5) edge [-,double equal sign distance] (m-4-5)\n (m-4-5) edge [-,double equal sign distance] (m-3-5)\n (m-3-5) edge [-,double equal sign distance] (m-2-5)\n (m-2-5) edge [-,double equal sign distance] (m-1-5)\n\n (m-4-4) edge [-,double equal sign distance] (m-3-4)\n (m-3-4) edge [-,double equal sign distance] (m-2-4)\n\n (m-3-3) edge [-,double equal sign distance] (m-2-3)\n\n (m-2-1) edge node [left] {\\footnotesize $0$} (m-1-1)\n (m-2-1) edge node [above] {\\footnotesize $\\begin{pmatrix} 1\\\\ 1\\end{pmatrix}$} (m-2-2)\n (m-2-2) edge node [right] {\\footnotesize $(1,-1)$} (m-1-2)\n (m-2-3) edge node [right] {\\footnotesize $(1,-1)$} (m-1-3)\n (m-2-4) edge node [right] {\\footnotesize $(1,-1)$} (m-1-4)\n\n (m-5-4) edge node [above] {\\footnotesize $0$} (m-5-5)\n (m-5-4) edge node [left] {\\footnotesize $\\begin{pmatrix} 1\\\\ 1\\end{pmatrix}$} (m-4-4)\n (m-4-4) edge node [above] {\\footnotesize $(1,-1)$} (m-4-5)\n (m-3-4) edge node [above] {\\footnotesize $(1,-1)$} (m-3-5)\n (m-2-4) edge node [above] {\\footnotesize $(1,-1)$} (m-2-5)\n\n (m-3-2) edge node [left] {\\footnotesize $\\begin{pmatrix} 1\\\\ 0\\end{pmatrix}$} (m-2-2)\n (m-3-2) edge node [above] {\\footnotesize $\\begin{pmatrix} 1\\\\ 0\\end{pmatrix}$} (m-3-3)\n\n (m-4-3) edge node [left] {\\footnotesize $\\begin{pmatrix} 0\\\\ 1\\end{pmatrix}$} (m-3-3)\n (m-4-3) edge node [above] {\\footnotesize $\\begin{pmatrix} 0\\\\ 1\\end{pmatrix}$} (m-4-4)\n\n (m-3-1) edge (m-2-1)\n (m-4-1) edge (m-3-1)\n (m-5-1) edge (m-4-1)\n\n (m-4-2) edge (m-3-2)\n (m-5-2) edge (m-4-2)\n\n (m-5-3) edge (m-4-3)\n\n (m-5-1) edge (m-5-2)\n (m-5-2) edge (m-5-3)\n (m-5-3) edge (m-5-4)\n\n (m-4-1) edge (m-4-2)\n (m-4-2) edge (m-4-3)\n \n (m-3-1) edge (m-3-2)\n ;\n \\end{tikzpicture}\n \\]\n\\end{definition}\n\n\n\n\\begin{restatable}{lemma}{Gindecomposable}\n \\label{lemma:G-is-indecomposable}\n The persistence module $\\Gsf$ is indecomposable.\n More specifically, the morphism of rings $\\End(\\Gsf) \\to \\End(\\Gsf(4,4)) \\cong \\kbb$ given by evaluation at $(4,4)$ is an isomorphism.\n\\end{restatable}\n\n\\begin{proof}\n\n\n Let $f : \\Gsf \\to \\Gsf$ be an endomorphism.\n If $f(4,4) : \\kbb \\to \\kbb$ is given by multiplication by $\\alpha$, then, for all $0 \\leq j \\leq 4$, we have that $f(4,j) : \\kbb \\to \\kbb$ and $f(j,4) : \\kbb \\to \\kbb$ must also be given by multiplication by $\\alpha$.\n\n Assume that $f(1,2) : \\kbb \\to \\kbb$ is given by multiplication by $\\beta$ and that $f(2,1) : \\kbb \\to \\kbb$ is given by multiplication by $\\gamma$.\n Since the structure morphisms $\\kbb = \\Gsf(1,2) \\to \\Gsf(4,4) = \\kbb$ is the identity, we must have $\\beta = \\alpha$.\n By an analogous argument, we have $\\gamma = \\alpha$.\n This also implies that for all $(j,k)$ with $\\Gsf(j,k) = \\kbb^2$ the endomorphism $f(j,k)$ is given by multiplication by $\\alpha$.\n \n \n To conclude, note that, by the definition of the structure morphisms $\\Gsf(0,3) \\to \\Gsf(1,3)$ and $\\Gsf(3,0) \\to \\Gsf(3,1)$, we have that $f(0,3)$ and $f(3,0)$ are also given by multiplication by $\\alpha$.\n Thus, $\\End(\\Gsf) \\cong \\End(\\Gsf(4,4)) \\cong \\kbb$, and thus $\\Gsf$ is indecomposable, by \\cref{lemma:indecomposable-local-ring}.\n\\end{proof}\n\nThe crucial property of $\\Gsf$ of importance to us, other than being indecomposable, is the fact that there are two pairs of adjacent copies of the ground field $\\kbb$ connected by a zero map, on the bottom right at $(0,3)$ and $(0,4)$, and on the top left at $(3,0)$ and $(4,0)$.\n\n\\begin{definition}\n \\label{definition:1-dim-corner}\n Let $X : \\Rbf^2 \\to \\mathbf{vec}$.\n We say that $X$ \\define{admits a thin corner} if there exists $\\epsilon > 0$, $X' : (\\epsilon \\Zbf)^2 \\to \\mathbf{vec}$, and $(x,y) \\in (\\epsilon \\Zbf)^2$ such that $X$ is isomorphic to the extension of $X'$, and we have $X'(x,y) = \\kbb$ and $X'(x-\\epsilon,y) = X'(x,y-\\epsilon) = 0$.\n\\end{definition}\n\nSee \\cref{figure:diagram-proof-main-lemma}(1.) for a schematic depiction of two modules admitting a thin corner.\n\n\\begin{lemma}\n \\label{lemma:add-1-dim-corner}\n Let $A : \\Rbf^2 \\to \\mathbf{vec}$ be finitely presentable, indecomposable, and isomorphic to the extension of a $\\Pscr$-persistence module for some regular grid $\\Pscr \\subseteq \\Rbf^2$.\n For every $\\delta > 0$, there exists $A_1 : \\Rbf^2 \\to \\mathbf{vec}$ indecomposable, finitely presentable, admitting a thin corner, isomorphic to an extension of a $\\Qscr$-persistence module for some regular grid $\\Qscr \\supseteq \\Pscr$, and such that $\\mathsf{im}\\left(\\eta^A_\\delta\\right) \\cong \\mathsf{im}\\big(\\eta^{A_1}_\\delta\\big)$.\n\\end{lemma}\n\\begin{proof}\n Assume that $\\Pscr = (\\epsilon \\Zbf)^2$.\n Let $m \\in \\Nbb$ be such that $\\epsilon\/m < \\delta$.\n Let $\\tau = \\epsilon\/m$.\n Since $A$ is finitely presentable, there exists $(x,y) \\in (\\tau\\Zbf)^2$ such that $A(x,y) \\neq 0$ and $A(x-\\tau,y) = 0 = A(x,y-\\tau)$.\n Thus, on the regular grid $((\\tau\/2)\\Zbf)^2$, and around index $(x,y)$ (highlighted entry), the module $A$ restricts as follows:\n \\[ \\footnotesize\n \\begin{tikzpicture}\n \\matrix (m) [matrix of math nodes,row sep=3em,column sep=1.5em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { A(x-\\tau,y+\\tau) & A(x-\\tau\/2,y+\\tau) & A(x,y+\\tau) \n & A(x+\\tau\/2,y+\\tau) & A(x+\\tau,y+\\tau) \\\\\n 0 & 0 & A(x,y+\\tau\/2)\n & A(x+\\tau\/2,y+\\tau\/2) & A(x+\\tau,y+\\tau\/2) \\\\\n 0 & 0 & |[fill=lightgray!25, rectangle, outer sep = 2pt, minimum size = 0]| A(x,y) & A(x+\\tau\/2,y) & A(x+\\tau,y) \\\\\n 0 & 0 & 0 & 0 & A(x+\\tau,y-\\tau\/2) \\\\\n 0 & 0 & 0 & 0 & A(x+\\tau,y-\\tau) \\\\};\n \\path[line width=0.5pt, -{>[width=4pt]}]\n (m-1-1) edge node [above] {$\\cong$} (m-1-2)\n (m-2-1) edge node [above] {$\\cong$} (m-2-2)\n (m-3-1) edge node [above] {$\\cong$} (m-3-2)\n (m-4-1) edge node [above] {$\\cong$} (m-4-2)\n (m-5-1) edge node [above] {$\\cong$} (m-5-2)\n\n (m-1-2) edge (m-1-3)\n (m-2-2) edge (m-2-3)\n (m-3-2) edge (m-3-3)\n (m-4-2) edge (m-4-3)\n (m-5-2) edge (m-5-3)\n\n (m-1-3) edge node [above] {$\\cong$} (m-1-4)\n (m-2-3) edge node [above] {$\\cong$} (m-2-4)\n (m-3-3) edge node [above] {$\\cong$} (m-3-4)\n (m-4-3) edge node [above] {$\\cong$} (m-4-4)\n (m-5-3) edge node [above] {$\\cong$} (m-5-4)\n\n (m-1-4) edge (m-1-5)\n (m-2-4) edge (m-2-5)\n (m-3-4) edge (m-3-5)\n (m-4-4) edge (m-4-5)\n (m-5-4) edge (m-5-5)\n\n (m-5-1) edge node [left] {$\\cong$} (m-4-1)\n (m-5-2) edge node [left] {$\\cong$} (m-4-2)\n (m-5-3) edge node [left] {$\\cong$} (m-4-3)\n (m-5-4) edge node [left] {$\\cong$} (m-4-4)\n (m-5-5) edge node [left] {$\\cong$} (m-4-5)\n\n (m-4-1) edge (m-3-1)\n (m-4-2) edge (m-3-2)\n (m-4-3) edge (m-3-3)\n (m-4-4) edge (m-3-4)\n (m-4-5) edge (m-3-5)\n\n (m-3-1) edge node [left] {$\\cong$} (m-2-1)\n (m-3-2) edge node [left] {$\\cong$} (m-2-2)\n (m-3-3) edge node [left] {$\\cong$} (m-2-3)\n (m-3-4) edge node [left] {$\\cong$} (m-2-4)\n (m-3-5) edge node [left] {$\\cong$} (m-2-5)\n\n (m-2-1) edge (m-1-1)\n (m-2-2) edge (m-1-2)\n (m-2-3) edge (m-1-3)\n (m-2-4) edge (m-1-4)\n (m-2-5) edge (m-1-5)\n ;\n \\end{tikzpicture}\n \\]\n where the isomorphisms are due to the fact that $A$ is an extension of a persistence module on the regular grid $(\\tau\\Zbf)^2$.\n\n Let $\\kbb \\to A(x,y)$ be any non-zero morphism.\n We now define a persistence module $A'$, and then define $A_1$ to be an extension of $A'$.\n Let $A' : ((\\tau\/2)\\Zbf)^2 \\to \\mathbf{vec}$ coincide with the restriction of $A$ to $((\\tau\/2)\\Zbf)^2$ at all places, except at $(x,y)$ where it takes the value $\\kbb$.\n The structure morphisms $A'(x,y) \\to A'(x,y+\\tau\/2) = A(x,y+\\tau\/2) $ and $A'(x,y) \\to A'(x+\\tau\/2, y) = A(x+\\tau\/2, y)$ are defined by composing the corresponding structure morphisms of $A$ with the map $\\kbb \\to A(x,y)$.\n Thus, around index $(x,y)$ (again highlighted), the module $A'$ looks as follows:\n \\[ \\footnotesize\n \\begin{tikzpicture}\n \\matrix (m) [matrix of math nodes,row sep=3em,column sep=1.5em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { A(x-\\tau,y+\\tau) & A(x-\\tau\/2,y+\\tau) & A(x,y+\\tau) \n & A(x+\\tau\/2,y+\\tau) & A(x+\\tau,y+\\tau) \\\\\n 0 & 0 & A(x,y+\\tau\/2)\n & A(x+\\tau\/2,y+\\tau\/2) & A(x+\\tau,y+\\tau\/2) \\\\\n 0 & 0 & |[fill=lightgray!25, rectangle, outer sep = 2pt, minimum size = 0]| \\kbb & A(x+\\tau\/2,y) & A(x+\\tau,y) \\\\\n 0 & 0 & 0 & 0 & A(x+\\tau,y-\\tau\/2) \\\\\n 0 & 0 & 0 & 0 & A(x+\\tau,y-\\tau) \\\\};\n \\path[line width=0.5pt, -{>[width=4pt]}]\n (m-1-1) edge (m-1-2)\n (m-2-1) edge (m-2-2)\n (m-3-1) edge (m-3-2)\n (m-4-1) edge (m-4-2)\n (m-5-1) edge (m-5-2)\n\n (m-1-2) edge (m-1-3)\n (m-2-2) edge (m-2-3)\n (m-3-2) edge (m-3-3)\n (m-4-2) edge (m-4-3)\n (m-5-2) edge (m-5-3)\n\n (m-1-3) edge (m-1-4)\n (m-2-3) edge (m-2-4)\n (m-3-3) edge (m-3-4)\n (m-4-3) edge (m-4-4)\n (m-5-3) edge (m-5-4)\n\n (m-1-4) edge (m-1-5)\n (m-2-4) edge (m-2-5)\n (m-3-4) edge (m-3-5)\n (m-4-4) edge (m-4-5)\n (m-5-4) edge (m-5-5)\n\n (m-5-1) edge (m-4-1)\n (m-5-2) edge (m-4-2)\n (m-5-3) edge (m-4-3)\n (m-5-4) edge (m-4-4)\n (m-5-5) edge (m-4-5)\n\n (m-4-1) edge (m-3-1)\n (m-4-2) edge (m-3-2)\n (m-4-3) edge (m-3-3)\n (m-4-4) edge (m-3-4)\n (m-4-5) edge (m-3-5)\n\n (m-3-1) edge (m-2-1)\n (m-3-2) edge (m-2-2)\n (m-3-3) edge (m-2-3)\n (m-3-4) edge (m-2-4)\n (m-3-5) edge (m-2-5)\n\n (m-2-1) edge (m-1-1)\n (m-2-2) edge (m-1-2)\n (m-2-3) edge (m-1-3)\n (m-2-4) edge (m-1-4)\n (m-2-5) edge (m-1-5)\n ;\n \\end{tikzpicture}\n \\]\n We have the following isomorphisms of endomorphism rings:\n \\[\n \\End(A) \\cong \\End\\left(A|_{((\\tau\/2)(2\\Zbf+1))^2}\\right) = \\End\\left(A'|_{((\\tau\/2)(2\\Zbf+1))^2}\\right) \\cong \\End\\left(A'|_{((\\tau\/2)\\Zbf)^2}\\right).\n \\]\n The first isomorphism is due to the fact that $A$ is isomorphic to the extension of a $(\\tau\\Zbf)^2$-persistence module.\n The equality is because\n $A|_{((\\tau\/2)(2\\Zbf+1))^2} = A'|_{((\\tau\/2)(2\\Zbf+1))^2}$.\n For the last isomorphism, note that the value of $A'$ at any element of the grid $((\\tau\/2)\\Zbf)^2$ other than $(x,y)$ is isomorphic, through a structure morphism of $A'$, to the value of $A'$ at an element of the grid $((\\tau\/2)(2\\Zbf+1))^2$, by construction.\n Moreover, by construction, the structure morphism $A'(x,y) = \\kbb \\to A'(x+\\tau\/2, y+\\tau\/2)$ is non-zero, and thus a monomorphism; and $(x+\\tau\/2, y+\\tau\/2) \\in ((\\tau\/2)(2\\Zbf+1))^2$.\n It follows that the action of an endomorphism of $A'$ is completely determined by the action of the endomorphism on the restriction $A'|_{((\\tau\/2)(2\\Zbf+1))^2}$, proving the last isomorphism.\n\n Thus, $A'$ is indecomposable by \\cref{lemma:indecomposable-local-ring} and the fact that $A$ is indecomposable.\n Define $A_1 = \\widehat{A'}$, the extension of $A'$ to $\\Rbf^2$.\n Note that $A_1$ is indecomposable, finitely presentable, and satisfies $\\mathsf{im}\\big(\\eta^{A_1}_\\delta\\big) = \\mathsf{im}\\left(\\eta^{A}_\\delta\\right)$.\n Moreover, $A_1$ admits a thin corner, by construction.\n\\end{proof}\n\n\\begin{definition}\n \\label{definition:horizontal-antenna-attachment}\n Let $X : \\Rbf^2 \\to \\mathbf{vec}$ and $(x,y) \\in \\Rbf^2$.\n We say that $X$ \\define{admits a horizontal antenna attachment} at $(x,y)$ if there exists $\\epsilon > 0$ and $X' : (\\epsilon \\Zbf)^2 \\to \\mathbf{vec}$, such that $(x,y) \\in (\\epsilon \\Zbf)^2$, $X$ is isomorphic to the extension of $X'$, and we have $X'(x,y) = \\kbb$, $X'(x-k\\epsilon,y) = 0$ for all $k \\geq 1 \\in \\Nbb$, and the structure morphism $X'(x,y) \\to X'(x,y+\\epsilon)$ is zero.\n\\end{definition}\n\nFor example, the extension $\\widehat\\Gsf: \\Rbf^2 \\to \\mathbf{vec}$ admits a horizontal antenna attachment.\nSee \\cref{figure:diagram-step-3} for a schematic depiction of two modules admitting a horizontal antenna attachment.\n\n\\begin{lemma}\n \\label{lemma:add-antenna-attachment}\n Let $A : \\Rbf^2 \\to \\mathbf{vec}$ be finitely presentable, indecomposable, and admitting a thin corner.\n For every $\\delta > 0$, there exists $A_1 : \\Rbf^2 \\to \\mathbf{vec}$ indecomposable, finitely presentable, admitting a horizontal antenna attachment, and such that $\\mathsf{im}\\left(\\eta^A_\\delta\\right) \\cong \\mathsf{im}\\big(\\eta^{A_1}_\\delta\\big)$.\n\\end{lemma}\n\n\\begin{proof}\n Since $A$ admits a thin corner, we may assume the following:\n there exists a regular grid $(\\epsilon \\Zbf)^2$ for some $\\epsilon > 0$ and $X : (\\epsilon \\Zbf)^2 \\to \\mathbf{vec}$ such that $A \\cong \\widehat{X}$, and such that there exists $(x,y) \\in (\\epsilon \\Zbf)^2$ with $X(x,y) = \\kbb$ and $X(x-\\epsilon,y) = X(x,y-\\epsilon) = 0$.\n\n Let $m \\in \\Nbb$ be such that $\\epsilon\/m < \\delta$ and let $\\tau = \\epsilon\/m$.\n We make use of the regular grid $((\\tau\/5)\\Zbf)^2$.\n By extending $X$ to $\\Rbf^2$ and restricting to $((\\tau\/5)\\Zbf)^2$, we get a module $X' : ((\\tau\/5)\\Zbf)^2 \\to \\mathbf{vec}$ that, when restricted to the finite grid\n \\begin{align*}\n \\Tscr = \\{(&x-\\tau\/5,x,x+\\tau\/5,x+2\\tau\/5,x+3\\tau\/5,x+4\\tau\/5)\\}\\\\\n &\\times \\{(y-\\tau\/5,y,y+\\tau\/5,y+2\\tau\/5,y+3\\tau\/5,y+4\\tau\/5)\\},\n \\end{align*}\n looks as follows (index $(x,y)$ highlighted):\n \\[\\footnotesize\n \\begin{tikzpicture}\n \\matrix (m) [matrix of math nodes,row sep=3em,column sep=3em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { 0 & \\kbb & \\kbb & \\kbb & \\kbb & \\kbb \\\\\n 0 & \\kbb & \\kbb & \\kbb & \\kbb & \\kbb \\\\\n 0 & \\kbb & \\kbb & \\kbb & \\kbb & \\kbb \\\\\n 0 & \\kbb & \\kbb & \\kbb & \\kbb & \\kbb \\\\\n 0 & |[fill=lightgray!25, rectangle, outer sep = 2pt, minimum size = 0]| \\kbb & \\kbb & \\kbb & \\kbb & \\kbb \\\\\n 0 & 0 & 0 & 0 & 0 & 0 \\\\};\n \\path[line width=0.5pt, -{>[width=6pt]}]\n (m-1-2) edge [-,double equal sign distance] (m-1-3)\n (m-1-3) edge [-,double equal sign distance] (m-1-4)\n (m-1-4) edge [-,double equal sign distance] (m-1-5)\n (m-1-5) edge [-,double equal sign distance] (m-1-6)\n (m-2-3) edge [-,double equal sign distance] (m-2-4)\n (m-2-4) edge [-,double equal sign distance] (m-2-5)\n (m-3-4) edge [-,double equal sign distance] (m-3-5)\n (m-5-6) edge [-,double equal sign distance] (m-4-6)\n (m-4-6) edge [-,double equal sign distance] (m-3-6)\n (m-3-6) edge [-,double equal sign distance] (m-2-6)\n (m-2-6) edge [-,double equal sign distance] (m-1-6)\n (m-4-5) edge [-,double equal sign distance] (m-3-5)\n (m-3-5) edge [-,double equal sign distance] (m-2-5)\n (m-3-4) edge [-,double equal sign distance] (m-2-4)\n (m-2-2) edge [-,double equal sign distance] (m-1-2)\n (m-2-2) edge [-,double equal sign distance] (m-2-3)\n (m-2-3) edge [-,double equal sign distance] (m-1-3)\n (m-2-4) edge [-,double equal sign distance] (m-1-4)\n (m-2-5) edge [-,double equal sign distance] (m-1-5)\n (m-5-5) edge [-,double equal sign distance] (m-5-6)\n (m-5-5) edge [-,double equal sign distance] (m-4-5)\n (m-4-5) edge [-,double equal sign distance] (m-4-6)\n (m-3-5) edge [-,double equal sign distance] (m-3-6)\n (m-2-5) edge [-,double equal sign distance] (m-2-6)\n (m-3-3) edge [-,double equal sign distance] (m-2-3)\n (m-3-3) edge [-,double equal sign distance] (m-3-4)\n (m-4-4) edge [-,double equal sign distance] (m-3-4)\n (m-4-4) edge [-,double equal sign distance] (m-4-5)\n (m-3-2) edge [-,double equal sign distance] (m-2-2)\n (m-4-2) edge [-,double equal sign distance] (m-3-2)\n (m-5-2) edge [-,double equal sign distance] (m-4-2)\n (m-4-3) edge [-,double equal sign distance] (m-3-3)\n (m-5-3) edge [-,double equal sign distance] (m-4-3)\n (m-5-4) edge [-,double equal sign distance] (m-4-4)\n (m-5-2) edge [-,double equal sign distance] (m-5-3)\n (m-5-3) edge [-,double equal sign distance] (m-5-4)\n (m-5-4) edge [-,double equal sign distance] (m-5-5)\n (m-4-2) edge [-,double equal sign distance] (m-4-3)\n (m-4-3) edge [-,double equal sign distance] (m-4-4)\n (m-3-2) edge [-,double equal sign distance] (m-3-3)\n\n (m-1-1) edge (m-1-2)\n (m-2-1) edge (m-2-2)\n (m-3-1) edge (m-3-2)\n (m-4-1) edge (m-4-2)\n (m-5-1) edge (m-5-2)\n (m-6-1) edge (m-6-2)\n\n (m-6-2) edge (m-6-3)\n (m-6-3) edge (m-6-4)\n (m-6-4) edge (m-6-5)\n (m-6-5) edge (m-6-6)\n\n (m-2-1) edge (m-1-1)\n (m-3-1) edge (m-2-1)\n (m-4-1) edge (m-3-1)\n (m-5-1) edge (m-4-1)\n (m-6-1) edge (m-5-1)\n\n (m-6-2) edge (m-5-2)\n (m-6-3) edge (m-5-3)\n (m-6-4) edge (m-5-4)\n (m-6-5) edge (m-5-5)\n (m-6-6) edge (m-5-6)\n ;\n \\end{tikzpicture}\n \\]\n Let $A' : ((\\tau\/5)\\Zbf)^2 \\to \\mathbf{vec}$ coincide with $X'$ at all places, except at the subgrid $\\Tscr$, where we use a copy of the module $\\Gsf$ as follows (index $(x,y)$ highlighted):\n \\[ \\footnotesize\n \\begin{tikzpicture}\n \\matrix (m) [matrix of math nodes,row sep=3em,column sep=3em,minimum width=2em,nodes={text height=1.75ex,text depth=0.25ex}]\n { 0 & \\kbb & \\kbb & \\kbb & \\kbb & \\kbb \\\\\n 0 & \\kbb & \\kbb^2 & \\kbb^2 & \\kbb^2 & \\kbb \\\\\n 0 & 0 & \\kbb & \\kbb^2 & \\kbb^2 & \\kbb \\\\\n 0 & 0 & 0 & \\kbb & \\kbb^2 & \\kbb \\\\\n 0 & |[fill=lightgray!25, rectangle, outer sep = 2pt, minimum size = 0]|0 & 0 & 0 & \\kbb & \\kbb \\\\\n 0 & 0 & 0 & 0 & 0 & 0 \\\\};\n \\path[line width=0.5pt, -{>[width=6pt]}]\n (m-1-2) edge [-,double equal sign distance] (m-1-3)\n (m-1-3) edge [-,double equal sign distance] (m-1-4)\n (m-1-4) edge [-,double equal sign distance] (m-1-5)\n (m-1-5) edge [-,double equal sign distance] (m-1-6)\n\n (m-2-3) edge [-,double equal sign distance] (m-2-4)\n (m-2-4) edge [-,double equal sign distance] (m-2-5)\n\n (m-3-4) edge [-,double equal sign distance] (m-3-5)\n\n (m-5-6) edge [-,double equal sign distance] (m-4-6)\n (m-4-6) edge [-,double equal sign distance] (m-3-6)\n (m-3-6) edge [-,double equal sign distance] (m-2-6)\n (m-2-6) edge [-,double equal sign distance] (m-1-6)\n\n (m-4-5) edge [-,double equal sign distance] (m-3-5)\n (m-3-5) edge [-,double equal sign distance] (m-2-5)\n\n (m-3-4) edge [-,double equal sign distance] (m-2-4)\n\n (m-2-2) edge node [left] {\\footnotesize $0$} (m-1-2)\n (m-2-2) edge node [above] {\\footnotesize $\\begin{pmatrix} 1\\\\ 1\\end{pmatrix}$} (m-2-3)\n (m-2-3) edge node [right] {\\footnotesize $(1,-1)$} (m-1-3)\n (m-2-4) edge node [right] {\\footnotesize $(1,-1)$} (m-1-4)\n (m-2-5) edge node [right] {\\footnotesize $(1,-1)$} (m-1-5)\n\n (m-5-5) edge node [above] {\\footnotesize $0$} (m-5-6)\n (m-5-5) edge node [left] {\\footnotesize $\\begin{pmatrix} 1\\\\ 1\\end{pmatrix}$} (m-4-5)\n (m-4-5) edge node [above] {\\footnotesize $(1,-1)$} (m-4-6)\n (m-3-5) edge node [above] {\\footnotesize $(1,-1)$} (m-3-6)\n (m-2-5) edge node [above] {\\footnotesize $(1,-1)$} (m-2-6)\n\n (m-3-3) edge node [left] {\\footnotesize $\\begin{pmatrix} 1\\\\ 0\\end{pmatrix}$} (m-2-3)\n (m-3-3) edge node [above] {\\footnotesize $\\begin{pmatrix} 1\\\\ 0\\end{pmatrix}$} (m-3-4)\n\n (m-4-4) edge node [left] {\\footnotesize $\\begin{pmatrix} 0\\\\ 1\\end{pmatrix}$} (m-3-4)\n (m-4-4) edge node [above] {\\footnotesize $\\begin{pmatrix} 0\\\\ 1\\end{pmatrix}$} (m-4-5)\n\n (m-3-2) edge (m-2-2)\n (m-4-2) edge (m-3-2)\n (m-5-2) edge (m-4-2)\n\n (m-4-3) edge (m-3-3)\n (m-5-3) edge (m-4-3)\n\n (m-5-4) edge (m-4-4)\n\n (m-5-2) edge (m-5-3)\n (m-5-3) edge (m-5-4)\n (m-5-4) edge (m-5-5)\n\n (m-4-2) edge (m-4-3)\n (m-4-3) edge (m-4-4)\n \n (m-3-2) edge (m-3-3)\n\n (m-1-1) edge (m-1-2)\n (m-2-1) edge (m-2-2)\n (m-3-1) edge (m-3-2)\n (m-4-1) edge (m-4-2)\n (m-5-1) edge (m-5-2)\n (m-6-1) edge (m-6-2)\n\n (m-6-2) edge (m-6-3)\n (m-6-3) edge (m-6-4)\n (m-6-4) edge (m-6-5)\n (m-6-5) edge (m-6-6)\n\n (m-2-1) edge (m-1-1)\n (m-3-1) edge (m-2-1)\n (m-4-1) edge (m-3-1)\n (m-5-1) edge (m-4-1)\n (m-6-1) edge (m-5-1)\n\n (m-6-2) edge (m-5-2)\n (m-6-3) edge (m-5-3)\n (m-6-4) edge (m-5-4)\n (m-6-5) edge (m-5-5)\n (m-6-6) edge (m-5-6)\n ;\n ;\n \\end{tikzpicture}\n \\]\n We have the following isomorphisms of endomorphism rings:\n \\[\n \\End\\left(X'\\right) \\cong \\End\\left(X'|_{((\\tau\/5)(5\\Zbf + 4))^2}\\right) = \\End\\left(A'|_{((\\tau\/5)(5\\Zbf + 4))^2}\\right) \\cong \\End\\left(A'|_{((\\tau\/5)\\Zbf)^2}\\right).\n \\]\n The first isomorphism is due to the fact that $X'$ is isomorphic to the restriction of an extension of an $(\\epsilon\\Zbf)^2$-persistence module, and $\\epsilon$ is an integer multiple of $\\tau$, by construction.\n The equality is because \n $X'|_{((\\tau\/5)(5\\Zbf + 4))^2} = A'|_{((\\tau\/5)(5\\Zbf + 4))^2}$, by construction.\n For the last isomorphism, note that the value of $A'$ at any element of the grid $((\\tau\/5)\\Zbf)^2$ not belonging to $\\Tscr$ is isomorphic, through a structure morphism of $A'$, to the value of $A'$ at an element of the grid $((\\tau\/5)(5\\Zbf+4))^2$, by construction.\n Moreover, the action of an endomorphism of $A'$ on $A'|_\\Tscr$ is determined by the action of the endomorphism on $A'(x+4\\tau\/5,y+4\\tau\/5)$, by \\cref{lemma:G-is-indecomposable}; and $(x+4\\tau\/5,y+4\\tau\/5) \\in ((\\tau\/5)(5\\Zbf+4))^2$.\n It follows that the action of an endomorphism of $A'$ is completely determined by the action of the endomorphism on the restriction $A'|_{((\\tau\/5)(5\\Zbf+4))^2}$, proving the last isomorphism.\n\n Thus, $A'$ is indecomposable by \\cref{lemma:indecomposable-local-ring} and the fact that $A$, and hence $X'$, is indecomposable.\n Define $A_1$ to be the extension of $A'$ to the full poset $\\Rbf^2$.\n Note that $A_1$ is indecomposable, finitely presentable, and such that $\\mathsf{im}\\left(\\eta^{A}_\\delta\\right) = \\mathsf{im}\\big(\\eta^{A_1}_\\delta\\big)$.\n Moreover, $A_1$ admits a horizontal antenna attachment at $(x,y+3\\tau\/5)$, by construction.\n\\end{proof}\n\n\\begin{lemma}\n \\label{lemma:sheaf-condition-persistence-module}\n Let $\\epsilon > 0$ and let $a \\in \\epsilon \\Zbf$.\n Let $s(a) = \\{(x,y) \\in (\\epsilon \\Zbf)^2 : x \\leq a\\}$, $g(a) = \\{(x,y) \\in (\\epsilon \\Zbf)^2 : x \\geq a\\}$, and $e(a) = \\{(x,y) \\in (\\epsilon \\Zbf)^2 : x = a\\}$, which are all subposets of $(\\epsilon \\Zbf)^2$.\n Assume given persistence modules $X : s(a) \\to \\mathbf{vec}$ and $Y : g(a) \\to \\mathbf{vec}$ such that $X|_{e(a)} = Y|_{e(a)}$.\n Then, there exists a unique persistence module $C : (\\epsilon\\Zbf)^2 \\to \\mathbf{vec}$ such that $C|_{s(a)} = X$ and $C|_{g(a)} = Y$.\n If, moreover, $X$ is indecomposable and, for every indecomposable summand $I$ of $Y$, we have that $I|_{e(a)} \\neq 0$, then $C$ is indecomposable.\n\\end{lemma}\n\\begin{proof}\n The statement about the existence and uniqueness of $C$ is clear.\n Now, assume that $C \\cong S \\oplus T$.\n On one hand, we have $X \\cong S|_{s(a)} \\oplus T|_{s(a)}$.\n Since $X$ is indecomposable, without loss of generality, it must be the case that $T_{g(a)} = 0$.\n In particular, $T_{e(a)} = 0$.\n On the other hand, $Y \\cong S|_{g(a)} \\oplus T|_{g(a)}$, and, since all the indecomposable summands $I$ of $Y$ satisfy $I|_{e(a)} \\neq 0$, we must have $T|_{g(a)} = 0$.\n Thus, $T = 0$ and $C$ does not admit any non-trivial decompositions.\n\\end{proof}\n\n\\begin{figure}[h]\n \\centering\n \\includegraphics[width=1\\linewidth]{pictures\/step-3.eps}\n \\caption{A schematic summary of the main constructions in the proof of \\cref{lemma:actual-tacking}.}\n \\label{figure:diagram-step-3}\n\\end{figure}\n\n\n\\begin{lemma}\n \\label{lemma:actual-tacking}\n Let $A,B : \\Rbf^2 \\to \\mathbf{vec}$ be finitely presentable indecomposable, and isomorphic to extensions of $\\Pscr$-persistence modules for some regular grid $\\Pscr \\subseteq \\Rbf^2$.\n Assume that $A$ and $B$ both admit horizontal antenna attachments at points in the grid $\\Pscr$.\n For every $\\delta > 0$, there exists a regular grid $\\Qscr \\supseteq \\Pscr$ and $M : \\Rbf^2 \\to \\mathbf{vec}$ with $M$ is indecomposable, finitely presentable, isomorphic to the extension of a $\\Qscr$-persistence module, and such that\n \\[\n \\mathsf{im}\\left(\\eta^M_\\delta\\right) \\cong \\mathsf{im}\\left(\\eta^A_\\delta\\right) \\oplus \\mathsf{im}\\left(\\eta^B_\\delta\\right).\n \\]\n\\end{lemma}\n\\begin{proof}\n Without loss of generality, we may assume that $A$ and $B$ are extensions of modules $A'$ and $B'$ defined over a regular grid $\\Qscr = (\\epsilon\\Zbf)^2$ with $\\epsilon < \\delta\/5$.\n Let $(x,y) \\in \\Qscr$ and $(w,z) \\in \\Qscr$ be horizontal antenna attachments for $A$ and $B$, respectively.\n Without loss of generality, we may assume that $y \\geq z$.\n Choose $a,b \\in \\epsilon\\Zbf$ with $a < \\min(x,w)$ and $b > y$.\n Let the subposets $s(a)$, $g(a)$, and $e(a)$ of $(\\epsilon\\Zbf)^2$ be as in \\cref{lemma:sheaf-condition-persistence-module}.\n\n We now use \\cref{lemma:sheaf-condition-persistence-module} to construct a persistence module $M$ as in the statement.\n In order to do this, we construct persistence modules $X$ and $Y$ over the grids $s(a)$ and $g(a)$, respectively.\n \\cref{figure:diagram-step-3} is a schematic summary of these constructions.\n\n Define a persistence module $Y_1 : g(a) \\to \\mathbf{vec}$ as follows:\n \\[\n Y_1(c,d) =\n \\begin{cases}\n \\kbb & \\text{if $d = y$ and $a \\leq c \\leq x$} \\\\\n A'(c,d) & \\text{else}\n \\end{cases}\n \\]\n The structure morphisms are taken to be those of $A'$ wherever that makes sense.\n The structure morphisms $\\kbb = Y_1(c,y) \\to Y_1(c+\\epsilon,y) = \\kbb$ for $a \\leq c \\leq x-\\epsilon$ are taken to be the identity, and the structure morphisms $Y_1(c,y) \\to Y_1(c,y+\\epsilon)$ and $Y_1(c,y-\\epsilon) \\to Y_1(c,y)$ are taken to be zero.\n It is straightforward to check that $Y_1$ is a well-defined persistence module and that $\\End(Y_1) \\cong \\End(A')$.\n Thus, $Y_1$ is indecomposable.\n Define $Y_2$ in an entirely analogous way using $B'$ and its horizontal antenna attachment.\n Let $Y = Y_1 \\oplus Y_2$.\n\n Consider the following subgrid of $\\epsilon \\Zbf$:\n \\[\n \\Tscr = \\{(a-5\\epsilon,a-4\\epsilon,a-3\\epsilon,a-2\\epsilon,a-\\epsilon)\\} \\times \\{(b, b+\\epsilon, b+2\\epsilon, b+3\\epsilon, b+4\\epsilon)\\}.\n \\]\n Let us assume that $y > z$; the case $y=z$ is similar.\n Define a persistence module $X : s(a) \\to \\mathbf{vec}$ as follows:\n \\[\n X(c,d) =\n \\begin{cases}\n \\Gsf\\left(\\frac{(c,d)-(a-5\\epsilon,b)}{\\epsilon}\\right) & \\text{if $(c,d) \\in \\Tscr$} \\\\\n \\kbb & \\text{if $c=a-\\epsilon$ and $y \\leq d \\leq b$} \\\\\n \\kbb & \\text{if $c=a-2\\epsilon$ and $z \\leq d \\leq b$} \\\\\n \\kbb & \\text{if $(c,d)=(a-\\epsilon,z)$} \\\\\n 0 & \\text{else}\n \\end{cases}\n \\]\n For the structure morphisms corresponding to pairs of elements of $\\Tscr$ we use the structure morphisms of $\\Gsf$; for the remaining vertical structure morphisms we use the identity of $\\kbb$ when possible and the zero morphism otherwise; having done this, for the remaining horizontal morphisms we use the identity of $\\kbb$ when possible (taking commutativity into account) and the zero morphism otherwise (in particular, all horizontal maps with second coordinate different from $y$ and $z$).\n \n It is again straightforward to check that $X$ is a well-defined persistence module and that $\\End(X) = \\End(\\Gsf)$.\n Thus, $X$ is indecomposable.\n\n Finally, define $C : (\\epsilon \\Zbf)^2 \\to \\mathbf{vec}$ using \\cref{lemma:sheaf-condition-persistence-module}; it follows that $C$ is indecomposable.\n Finally, extend $C$ to the poset $\\Rbf^2$ to get a module $M$, which satisfies the conditions in the statement by construction.\n\\end{proof}\n\n\n\n\n\\begin{proof}[Proof of \\cref{lemma:main-lemma-attaching}]\n Let $A$ and $B$ be as in the statement.\n Using \\cref{lemma:add-1-dim-corner}, we may assume that $A$ and $B$ admit a thin corner, and using \\cref{lemma:add-antenna-attachment}, we may assume that $A$ and $B$ admit horizontal antenna attachments.\n \\cref{lemma:actual-tacking} now finishes the proof.\n\\end{proof}\n\n\\ifarxivversion\n\\bibliographystyle{alpha}\n\\fi\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} {"text":"\\section{introduction}\\label{sec_1}\nUltrafast pump-probe spectroscopy is a good tool to investigate the nonequilibrium properties of a given system since a pump pulse triggers ultrafast processes and a subsequent probe pulse monitors the pump-induced dynamical processes~\\cite{Mukamel1995, Diels1996, Krausz2009, Giannetti2016}.\nEspecially, by using femtosecond pulses, nonequilibrium dynamics of electrons can be detected since the timescale of the motion of electrons is of the order of a femtosecond.\nHowever, increasing the resolution of optical measurements in both the time and energy domains is difficult and limited by the uncertainty principle.\n\nRecently, ultrafast spectroscopic techniques have been advanced by using a transform-limited pulse, i.e., a pulse that has the minimum possible duration for a given spectral bandwidth, and have opened a new door to make both temporal and spectral resolutions as high as possible~\\cite{Diels1996}.\nThese techniques can disclose new ultrafast nonequilibrium phenomena.\nIn fact, by applying these techniques, interference in the energy domain has been observed in atomic systems and nanometric tips~\\cite{Wollenhaupt2002, Lindner2005, Kiffner2006, Milosevic2006, Kruger2018}.\nThis interference is applied to control the atomic storage medium for recording the information of optical pulses~\\cite{Leung1982, Carlson1983, Hemmer1994, Fleischauer2002, Ohmori2009}.\nHowever, as far as we know, there has been no such report on transient interference of pump-probe spectroscopy of band and Mott insulators both experimentally and theoretically.\n\nIn this paper, we investigate ultrafast pump-probe spectroscopy of band and Mott insulators and propose transient interference between temporary well-separated pulses in electron systems as in the case of atomic systems.\nWe formulate such transient interference in pump-probe spectroscopy of a two-band model.\nWe find that the existence of a continuum structure in the excitation spectrum is important for generating the transient interference since the continuum structure acts as a medium for storing the spectral information of the pump pulse and for creating interference between temporary well-separated pump and probe photons.\nThe information persists due to a memory effect, i.e., a relaxation process of electron systems.\nAs a result, the time-domain pump-probe spectrum depends on both probe energy $\\omega$ and the central frequency of the pump and probe pulses $\\Omega$ and thus oscillates with a frequency\n\\begin{align}\n\\label{probe_dep}\n\\omega_0 = \\omega - \\Omega.\n\\end{align}\nIn order to demonstrate the transient interference in the presence of electron correlation, we perform numerical calculations of the pump-probe spectrum in a one-dimensional (1D) half-filled Hubbard model.\nMoreover, we find that bosons coupled to electrons in the two-band model make an additional contribution to the interference.\nBased on the result, we speculate that the transient interference will be observed in Mott insulators strongly correlated to magnons.\nFor the observation of the proposed transient interference, high resolution of measurements of both time and energy is required in ultrafast pump-probe spectroscopy.\nRecently, oscillations of electronic states above the charge-transfer gap in a two-dimensional (2D) Mott insulator Nd$_2$CuO$_4$ were observed on the reflectivity changes detected by pump-probe measurement with ultrashort pulses~\\cite{Miyamoto2018}. \nThe time and energy resolutions of the measurement are as high as $10 \\text{fs}$ and $0.01 \\text{eV}$, respectively.\nBy extracting the oscillatory components from the pump-probe spectrum, the oscillation component with the frequency indicated by Eq.~(\\ref{probe_dep}) was found~\\cite{Miyamoto2018}.\nWe propose that the transient interference will be one of the possible origins of the observed oscillations.\n\nThis paper is organized as follows. We introduce a two-band model, which is a minimal model to describe the interference effect by two photon pulses through an electron system, and show the pump-probe absorption spectrum in Sec.~\\ref{sec_2}. In Sec.~\\ref{sec_3}, we calculate the time-dependent optical conductivity at half filling just after pumping. The effect of bosons coupled to electrons on the pump-probe spectrum is discussed in Sec.~\\ref{sec_4}. Finally, a summary is given in Sec.~\\ref{sec_5}.\n\n\\section{Two-band model}\\label{sec_2}\nWe first introduce a two-band model, which is the minimal model to describe the interference effect by two photon pulses through an electron system, and analytically calculate the pump-probe absorption spectrum.\nWith the assumption of dipole transitions, the Hamiltonian of the two-band model under the time-dependent electric field reads\n\\begin{align*}\n\\mathcal{H} =& \\sum_{\\bm{k}} \\varepsilon_{\\bm{k}} c_{\\text{c}\\bm{k}}^\\dag c_{\\text{c}\\bm{k}} + \\sum_{\\bm{k}} \\varrho_{\\bm{k}} c_{\\text{v}\\bm{k}}^\\dag c_{\\text{v}\\bm{k}} \\nonumber\\\\\n&-\\sum_{\\bm{k}} \\left( d_{\\text{cv}} \\mathcal{E}(t) c_{\\text{c}\\bm{k}}^\\dag c_{\\text{v}\\bm{k}} + d_{\\text{cv}}^* \\mathcal{E}(t) c_{\\text{v}\\bm{k}}^\\dag c_{\\text{c}\\bm{k}} \\right), \\nonumber\n\\end{align*}\nwhere $c_{\\text{c(v)}\\bm{k}}$ is an annihilation operator for fermions in the conduction (valence) band with momentum $\\bm{k}$. The energies of the conduction and valence bands are $\\varepsilon_{\\bm{k}} = \\varepsilon + \\frac{\\hbar ^2 \\bm{k}^2}{2m_{\\text{c}}}$ and $ \\varrho_{\\bm{k}} = \\varrho + \\frac{\\hbar ^2 \\bm{k}^2}{2m_{\\text{v}}}$,\nwhere $\\varepsilon$ and $\\varrho$ are the minimum and maximum of the conduction and valence bands, respectively, and $m_c$ and $m_v$ are the effective masses of electrons in the conduction and valence bands, respectively.\nWe introduce the interband dipole matrix element $d_{\\text{cv}}$ and external electric field $\\mathcal{E}(t)$. Hereafter, we set $\\hbar=1$.\n\nBy taking the long-wave-length limit of the electric field, the optical Bloch equation is written as~\\cite{HaugKoch}\n\\begin{align} \\label{Eq_bloch_eq_1}\n&\\left( \\frac{\\partial}{\\partial t} + i\\{ \\varepsilon _{\\bm{k}}-\\varrho_{\\bm{k}}-i \\gamma \\} \\right) p_{\\text{vc}}^0(\\bm{k},t) = d_{\\text{cv}} \\mathcal{E}(t) \\{ 1-2f_c(\\bm{k}) \\}\n\\end{align}\nand\n\\begin{align} \\label{Eq_bloch_eq_2}\n\\left( \\frac{\\partial}{\\partial t} + \\Gamma \\right) f_{\\text{c}}(\\bm{k},t) = -2\\text{Im}\\left[ d_{\\text{cv}} \\mathcal{E}(t) p_{\\text{vc}}^{0*}(\\bm{k},t) \\right],\n\\end{align}\nwhere $f_{\\text{c}}(\\bm{k}) = \\langle c_{\\text{c}\\bm{k}}^\\dag c_{\\text{c}\\bm{k}} \\rangle$ and $p_{\\text{vc}}^0(\\bm{k}) = \\langle c_{\\text{v}\\bm{k}}^\\dag c_{\\text{c}\\bm{k}} \\rangle$, with $\\langle \\cdots \\rangle$ representing the expectation value.\nWe introduce a phenomenological damping rate $\\Gamma$ for $f_{\\text{c}}$ and dephasing rate $\\gamma$ for $p_{\\text{vc}}^0$.\nWe consider an electric field $\\mathcal{E}(t) = \\frac{1}{2}\\left(\\mathcal{\\tilde{E}}(t)e^{-i\\Omega t} + \\mathcal{\\tilde{E}}^*(t)e^{i\\Omega t}\\right)$, where $\\mathcal{\\tilde{E}}(t)=2\\left\\{ \\mathcal{\\tilde{E}}_{\\text{p}}(t)e^{i\\bm{k}_p\\cdot \\bm{r}} + \\mathcal{\\tilde{E}}_{\\text{t}}(t)e^{i\\bm{k}_t\\cdot \\bm{r}} \\right\\}$, and the electric field and wave vector of the pump (probe) pulse are $\\mathcal{\\tilde{E}}_{\\text{p}}$ and $\\bm{k}_{\\text{p}}$ ($\\mathcal{\\tilde{E}}_{\\text{t}}$ and $\\bm{k}_{\\text{t}}$), respectively.\nIntroducing an expansion parameter $\\lambda$ through $\\mathcal{E}(t) \\rightarrow \\lambda \\mathcal{E}(t)$, we obtain $p_{\\text{vc}}^0 = \\lambda p_{\\text{vc}}^{0(1)} + \\lambda^2 p_{\\text{vc}}^{0(2)} + \\lambda^3 p_{\\text{vc}}^{0(3)} + \\cdots,\\; f_{\\text{c}} = \\lambda f_{\\text{c}}^{(1)} + \\lambda^2 f_{\\text{c}}^{(2)} + \\lambda^3 f_{\\text{c}}^{(3)} + \\cdots$.\nThe shape of the probe pulse is represented by the delta function $\\mathcal{\\tilde{E}}_{\\text{t}}(t) = \\mathcal{\\tilde{E}}_{\\text{t}} \\delta (t -\\tau)$ $(\\tau >0)$, where $\\tau$ is the delay time between the pump and probe pulses.\nThe pump-induced absorption change is given by $\\alpha = -\\textrm{Im}\\left[ d_{cv}^*\\chi(\\bm{k},\\omega) \\right].$\nTaking $\\mathcal{\\tilde{E}}_{\\text{p}}(t)=\\mathcal{\\tilde{E}}_{\\text{p}}e^{-\\sigma|t|}$ and with the rotating-wave approximation, the probe susceptibility is given by (see Appendix A)\n\\begin{align} \\label{chi0}\n&\\chi (\\bm{k},\\omega)\\simeq \\frac{p_{\\text{vc}}^{0(3)}(\\bm{k},\\omega)}{\\mathcal{E}_{\\text{t}}(\\omega)}\\nonumber \\\\\n&=\\frac{8d_{\\text{cv}}\\left|d_{\\text{cv}}\\right| {}^2 \\left| \\mathcal{\\tilde{E}}_p\\right| {}^2 e^{-\\left(\\sigma - \\gamma \\right)\\tau} e^{i \\tau \\left( - \\Omega +\\varepsilon _k-\\varrho _k\\right)} \\Gamma \\sigma }{\\left(i \\gamma +\\omega -\\varepsilon _k+\\varrho _k\\right)(i\\Gamma +i\\sigma + \\omega -\\Omega ) v_{\\bm{k}}^+ u_{\\bm{k}}^+ u_{\\bm{k}}^-}+\\cdots,\n\\end{align}\nwhere $u_{\\bm{k}}^{\\pm}=i\\gamma \\pm i\\sigma + \\Omega - \\varepsilon _{\\bm{k}} + \\varrho _{\\bm{k}}$ and $v_{\\bm{k}}^+ = i\\gamma +i\\Gamma -i\\sigma + \\Omega - \\varepsilon _{\\bm{k}}+ \\varrho _{\\bm{k}}$. \nIn the limit $\\gamma\\rightarrow 0$, the pole of the energy denominator $\\omega = \\varepsilon_{\\bm{k}} - \\varrho _{\\bm{k}}$ in the third term of $\\chi (\\bm{k},\\omega)$ gives rise to an oscillatory behavior of $e^{i(\\omega -\\Omega)\\tau}$ with decay $e^{-(\\sigma-\\gamma)\\tau}$.\nThis is the oscillation component indicated by Eq.~(\\ref{probe_dep}).\nSince the timescale where the oscillation persists is on the order of $\\gamma^{-1}$, real-time ultrafast dynamics should be observed with high accuracy~\\cite{Rhodes2013}.\n\nIn order to maintain the oscillation in the two-band model, we have to select a proper set of parameters that leads to the coherence and memory effect in the energy domain.\nFirst of all, we examine the coherence in the energy domain.\nWhen $\\sigma \\gg 1\/\\tau$, i.e., the pulse duration is much shorter than the time delay $\\tau$, we obtain $\\Delta t \\sim 0$, where $\\Delta t$ is the uncertainty in the time domain.\nSimultaneously, the energy uncertainty $\\Delta E \\sim \\infty$, leading to low energy resolution.\nAs a result, the interference in the energy domain is invisible.\nThis corresponds to the fact that the interference pattern vanishes in Young's double-slit experiment if the path of light is measured~\\cite{Englert1996, Durr1998}.\nIn fact, if the electric field of the pump pulse is represented by the $\\delta$ function, $p_{vc}^{0(3)}(\\bm{k},\\omega)$ completely cancels out $\\mathcal{E}_t(\\omega)$, which means that $\\chi(\\bm{k},\\omega)$ does not have the interference term $e^{i(\\omega -\\Omega)\\tau}$ (see Appendix A).\nIn contrast, when $\\sigma \\lesssim 1\/\\tau$, the coherence in the energy domain is obtained, which leads to the interference in energy space.\n\nSecond, we examine the memory effect.\nWhen $\\sigma \\ll \\gamma$, i.e., the pulse duration is longer than the dephasing time, $\\Delta t \\sim \\infty$ and $\\Delta E\\sim 0$ are simultaneously obtained.\nThis leads to the relaxation that holds true as long as electrons have well-defined energies, and their energy changes are slow with the timescale of $1\/\\Delta \\epsilon$, where $\\Delta \\epsilon$ is the characteristic energy exchange in a scattering event~\\cite{Aihara1982, Kuznetsov1991, Kuznetsov1991_2, Rossi2002, SchaeferWegener}.\nWhen $\\sigma \\gtrsim \\gamma$, the relaxation involving electrons with ill-defined energies starts to contribute to the memory effect.\nTherefore, if $\\sigma$ and $1\/\\tau$ are carefully controlled to realize $1\/\\tau \\gtrsim \\sigma \\gtrsim \\gamma$, both coherence in the energy domain and the memory effect are relevant, and the interference in the energy domain is maintained for the time $\\gamma^{-1}$.\nUsually, $\\gamma$ of a given system cannot be changed. However, if we make use of the quantum Zeno effect~\\cite{Misra1977, Itano1990, Kaulakys1997, Streed2006}, we might be able to control $\\gamma$, which can help us to observe our finding.\n\n\\section{Hubbard model}\\label{sec_3}\nPump-probe spectroscopy has been performed in strongly correlated systems to investigate exotic phenomena~\\cite{Okamoto2010, Okamoto2011, Filippis2012, Matsueda2012, Zala2013, Golez2014, Eckstein2014, Novelli2014, Prelovsek2015, Giannetti2016, Bittner2017, Bittner2018, Miyamoto2018}.\nEven in correlated electron systems, there is a continuum structure in the excitation spectrum. This indicates that interference effects similar to those in the two-band model may be realized, which will be demonstrated by using a 1D half-filled Hubbard model, which is given by\n\\begin{align}\nH=-t_\\mathrm{h}\\sum_{i,\\sigma} \\left( c^\\dagger_{i,\\sigma} c_{i+1,\\sigma} + \\mathrm{H.c.}\\right) + U\\sum_i n_{i,\\uparrow}n_{i,\\downarrow},\n\\label{H}\n\\end{align}\nwhere $c^\\dagger_{i\\sigma}$ is the creation operator of an electron with spin $\\sigma$ at site $i$, $n_{i,\\sigma}=c^\\dagger_{i,\\sigma}c_{i,\\sigma}$, $n_i=\\sum_\\sigma n_{i,\\sigma}$, and $t_\\mathrm{h}$ and $U$ are the nearest-neighbor hopping and on-site Coulomb interaction, respectively. Taking $t_\\mathrm{h}$ to be the unit of energy ($t_\\mathrm{h}=1$), we use $U=10$.\n\nWe investigate the probe-energy dependence of the optical conductivity of a Hubbard open chain with $L=10$, where $L$ is the number of sites.\nWe assume that both the pulses have the same shape of the vector potential given by $A(t)=A_0 e^{-(t-t_0)^2\/(2t_\\mathrm{d}^2)} \\cos[\\Omega(t-t_0)]$.\nWe set $A_0=0.1$, $t_0=3.0$, $t_\\mathrm{d}=0.5$, and $\\Omega=E_g =7.1$ for the pump pulse and $A_0=0.001$, $t_0=\\tau+3.0$, $t_\\mathrm{d}=0.02$, and $\\Omega=E_g =7.1$ for the probe pulse unless otherwise specified, where $E_g$ is the energy of the Mott gap.\nAn external spatially homogeneous electric field applied along the chain in the Hamiltonian can be incorporated via the Peierls substitution in the hopping terms as $c_{i,\\sigma}^\\dag c_{i+1,\\sigma} \\rightarrow e^{iA(t)}c_{i,\\sigma}^\\dag c_{i+1,\\sigma}$.\nUsing the method discussed in Refs.~\\cite{Lu2015, Shao2016}, we obtain the optical conductivity in the nonequilibrium system, $\\sigma(\\omega,\\tau) = \\frac{j_\\text{probe}(\\omega,\\tau) }{i(\\omega +i\\eta)LA_\\text{probe}(\\omega)}$, where $j_\\text{probe}(\\omega,\\tau)$ is the Fourier transform of the current induced by the probe pulse and $A_{\\text{probe}}(\\omega)$ is the Fourier transform of the vector potential of the probe pulse (see Appendix B for details).\n\nTo trace the temporal evolution of the system, we employ the time-dependent Lanczos method to evaluate $|\\psi (t)\\rangle$. \nHere $|\\psi(t+\\delta{t})\\rangle\\simeq\\sum_{l=1}^{M}{e^{-i\\epsilon_l\\delta{t}}}|\\phi_l\\rangle\\langle\\phi_l|\\psi(t)\\rangle$, where $\\epsilon_l$ and $|\\phi_l\\rangle$ are eigenvalues and eigenvectors of the tridiagonal matrix generated in the Lanczos iteration, respectively, $M$ is the dimension of the Lanczos basis, and $\\delta t$ is the minimum time step. We set $M=50$ and $\\delta t=0.02$.\n\n\\begin{figure}[t]\n \\centering\n \\includegraphics[clip, width=20pc]{Fig_OC_L10p_wpump10_wprobe_t_high.eps}\n \\caption{$\\text{Re} \\sigma (\\omega,\\tau)$ in the 1D half-filled Hubbard chain with $L=10$ and $U=10$, before pumping ($\\tau<0$) and after pumping ($\\tau=10$,\\;20,\\;30, and 40). Since the system is weakly excited, the dashed line for $\\tau<0$ is almost overlapped by the solid lines above $\\omega =7$.}\n \\label{band}\n\\end{figure}\n\nFigure~\\ref{band} shows the real part of the time-dependent optical conductivity $\\text{Re} \\sigma(\\omega,\\tau)$ of the Hubbard model.\nPhotoinduced decreases in the spectral weights at absorption peaks above the Mott gap are small since the system is weakly excited. \nThe pump photon excites carriers into an optically allowed odd-parity state.\nThe probe pulse couples in part to the odd-parity state, resulting in an excitation from the optically allowed state to an optically forbidden even-parity state.\nIn 1D Mott insulators with open boundary conditions, the optically forbidden state is located slightly above the optically allowed state~\\cite{Mizuno2000}.\nLow-energy in-gap excitation comes from the excitation from the optically allowed to forbidden state~\\cite{Lu2015}.\nInside the Mott gap, we find photoinduced low-energy spectral weights at $\\omega \\simeq1.2$, 2.2, and 3.3.\nThese energies correspond to the energy differences between the optically allowed populated state at $\\omega=7.1$ and the optically forbidden states.\n\nFigures~\\ref{probe_energy_dep_w0}(a)-\\ref{probe_energy_dep_w0}(e) show the $\\tau$ dependence of $\\text{Re} \\sigma(\\omega,\\tau)$ above the Mott gap with probe energy $\\omega=7.10$, $7.92$, $8.98$, $10.08$, and $11.18$, respectively, whose energies agree with the peak energies of the absorption spectrum in Fig.~\\ref{band}.\nWe find that the frequencies of the oscillations depend on $\\omega$.\nThe larger $\\omega$ is, the larger the frequency is, which is consistent with our argument in the two-band model discussed above.\n\\begin{figure}[t]\n \\centering\n \\includegraphics[clip, width=20pc]{OC_power.eps}\n \\caption{$\\text{Re} \\sigma(\\omega,\\tau)$ in the 1D half-filled Hubbard chain with $L=10$ and $U=10$ for (a) $\\omega=7.1$, (b) $\\omega=7.92$, (c) $\\omega=8.98$, (d) $\\omega=10.08$, and (e) $\\omega=11.18$. The power spectra of $\\text{Re} \\sigma (\\omega,\\tau)$ for (f) $\\omega=7.1$, (g) $\\omega=7.92$, (h) $\\omega=8.98$, (i) $\\omega=10.08$, and (j) $\\omega=11.18$.}\n \\label{probe_energy_dep_w0}\n\\end{figure}\n\nIn order to further examine the probe-energy dependence, we show the power spectra of $\\text{Re} \\sigma(\\omega,\\tau)$ with respect to $\\tau$ in Figs.~\\ref{probe_energy_dep_w0}(f)-\\ref{probe_energy_dep_w0}(j) for $\\omega=7.1$, 7.92, 8.98, 10.08, and 11.18, respectively.\nWe discuss two possible contributions to the power spectra.\nThe first one is the contribution from the Rabi oscillation, whose frequencies are related to the low-energy in-gap states at $\\omega=$1.2, 2.2, and 3.3.\nIn fact, we find the Rabi-oscillation contributions to the spectral weights at $\\omega_0=$1.2, 2.2, and 3.3 in Figs.~\\ref{probe_energy_dep_w0}(f)-\\ref{probe_energy_dep_w0}(j).\nSince our system is of finite size, energy levels are discretized.\nTherefore, there are oscillations with resonant frequencies between the levels.\nIn the thermodynamic limit, the number of the levels is infinite, and thus we expect that the contributions from a huge number of such resonances with various frequencies cancel out, giving rise to an inifinite number of infinitesimal weights in the power spectra.\nThus, we consider that the Rabi-oscillation contribution to the power spectra is visible only in finite-size systems and negligible in the thermodynamic limit.\n\nThe second one is the contribution from the interference effect, which gives rise to the $\\omega$ dependence of the pump-probe spectra as discussed in the two-band model. \nThe oscillations with the frequencies $\\omega-\\Omega$ appear at $\\omega_0=7.92-7.10=0.82$, $8.98-7.10=1.88$, $10.08-7.10=2.98$, and $11.18-7.10=4.08$.\nThese energies correspond to the energy difference between the levels at $\\omega=\\Omega=7.1$ and the excited states above the Mott gap, all of which belong to the same electronic states with odd parity.\nWe consider that this origin makes the dominant contribution to the power spectra in the thermodynamic limit.\nIn order to induce the transient interference, we should use the pump pulse whose spectrum covers some energy levels. Then we can store the information of the pump pulse in electronic states with a wide range of energies above the Mott gap.\n\nAccording to the two possible contributions to the power spectra, in Fig.~\\ref{probe_energy_dep_w0}(g), for example, we find peak structures at $\\omega_0=0.82$, 1.2, and 2.2.\nThe peak structures at $\\omega_0=1.2$ and 2.2 come from the Rabi oscillation of the two odd- and even-parity states.\nOn the other hand, the origin of the structure at $\\omega_0=0.82$ is the interference because $\\omega_0=0.82$ corresponds to one of the energy differences between the odd-odd states mentioned above.\nSimilarly, Figs.~\\ref{probe_energy_dep_w0}(h)-\\ref{probe_energy_dep_w0}(j) are understood in the same way (see Appendix B for details).\n\n\\section{electron-boson coupling in the two-band model} \\label{sec_4}\n\nFinally, we discuss the effect of bosons coupled to electrons on the probe-energy-dependent oscillation.\nNonequilibrium electron dynamics coupled to a boson driven by a laser has been extensively studied.\nFurthermore, since non-Markovian relaxation is important in electron systems coupled to a bosonic environment, open quantum systems with non-Markovian properties have been studied for a long time~\\cite{Jaynes1963, Caldeira1985, Leggett1987, BreuerPetruccione, Zurek2003, Reiter2014, Seetharam2015, Nazir2016, deVega2017}.\nThe additional Hamiltonian due to boson degrees of freedom is \n\\begin{align}\\label{H_ph}\n\\mathcal{H}_{\\text{ph}} = \\sum_{\\bm{q}} \\omega_{\\bm{q}} a_{\\bm{q}}^\\dag a_{\\bm{q}} + \\sum_{\\bm{k},\\bm{q}} g_{\\bm{q}} (a_{-\\bm{q}}^\\dag + a_{\\bm{q}}) (c_{\\text{c}\\bm{k+q}}^\\dag c_{\\text{c}\\bm{k}} + c_{\\text{v}\\bm{k+q}}^\\dag c_{\\text{v}\\bm{q}}),\n\\end{align}\nwhere $a_{\\bm{q}}$ is an annihilation operator for bosons with momentum $\\bm{q}$, $\\omega_{\\bm{q}}$ is the boson frequency, and $g_{\\bm{q}}$ is an electron-boson coupling constant.\n\nWe examine the two-band model with electron-boson coupling under the application of the exponential pump pulse.\nTotal polarization is given by $p_{\\text{vc}}^{}(\\bm{k},t)= p_{\\text{vc}}^{0}(\\bm{k},t) + p_{\\text{vc}}^{b}(\\bm{k},t)$, where $p_{\\text{vc}}^{0}(\\bm{k},t)$ is from the one-particle contribution, as discussed above, and $p_{\\text{vc}}^{b}(\\bm{k},t)$ is from the electron-boson coupling.\nSolving the kinetic equation with $\\mathcal{H}_{\\text{ph}}$ (see Appendix A), the probe susceptibility is given by\n\\begin{align} \\label{Eq_chi_b_main}\n&\\chi ^{b}(\\bm{k},\\omega)\\simeq \\frac{p_{\\text{vc}}^{b(3)}(\\bm{k},\\omega)}{\\mathcal{E}_{\\text{t}}(\\omega)}\\nonumber =\\sum _{\\bm{q}} g_{\\bm{q}}^2\\mathcal{N}_{\\bm{q}} \\cdot 4i\\sigma d_{\\text{cv}}\\left| d_{\\text{cv}}\\right|{}^2 \\left| \\mathcal{\\tilde{E}}_p\\right| {}^2 \\nonumber \\\\\n&\\times \\Biggl[\n\\frac{e^{-\\tau \\left(\\sigma-\\gamma\\right)} e^{i \\tau \\left( -\\Omega +\\varepsilon_{\\bm{k}} -\\varrho _{\\bm{k}}\\right)} \\left(-i\\gamma -2i \\Gamma - \\omega + \\varepsilon _{\\bm{k}}- \\varrho _{\\bm{k}}\\right) }{\\left(i \\gamma +\\omega -\\varepsilon _{\\bm{k}}+\\varrho _{\\bm{k}}\\right){}^2 \\left(i \\gamma +\\omega -\\varepsilon _{\\bm{k+q}}+\\varrho _{\\bm{k}}+\\omega _{\\bm{q}}\\right) v_{\\bm{k}}^+}\\nonumber \\\\\n&\\times \\frac{\\left(2 i \\gamma +2 \\omega -\\varepsilon _{\\bm{k}}-\\varepsilon _{\\bm{k+q}}+\\varrho _{\\bm{k}}+\\varrho _{\\bm{k-q}}+2 \\omega _{\\bm{q}}\\right)}{(i\\Gamma +i\\sigma + \\omega -\\Omega ) \\left(i \\gamma +\\omega -\\varepsilon _{\\bm{k}}+\\varrho _{\\bm{k-q}}+\\omega _{\\bm{q}}\\right) u_{\\bm{k}}^+u_{\\bm{k}}^-}\\Biggr]\\nonumber \\\\\n&+\\cdots,\n\\end{align}\nwhere $\\mathcal{N}_{\\bm{q}}=\\frac{1}{e^{\\omega_{\\bm{q}}\/k_BT}-1}$.\nIn the limit $\\gamma \\rightarrow 0$, the pole of the energy denominator $\\omega = \\varepsilon_{\\bm{k}} - \\varrho _{\\bm{k}}$ gives rise to an oscillatory behavior of $e^{i(\\omega -\\Omega)\\tau}$ with decay $e^{-(\\sigma -\\gamma)\\tau}$, which is the same behavior as the third term in Eq.~(\\ref{chi0}).\nTherefore, the information of pump and probe pulses is transmitted with the help of boson-assisted electron scattering, which gives one of the possible origins of the transient interference.\n\nIn Mott insulators, magnons are strongly coupled to photo-excited electrons in 2D Mott insulators, in contrast to the 1D Mott insulator where spin and charge degrees of freedom are separated.\nTherefore, the interference proposed in this work will be easily realized in the 2D Mott insulators.\nWe thus speculate that the oscillations observed by the pump-probe spectroscopy of the 2D Mott insulator Nd$_2$CuO$_4$~\\cite{Miyamoto2018} come from the interference effect. \nIn order to confirm this speculation, we need to investigate theoretically the pump-probe spectrum of the 2D half-filled Hubbard model, but it remains for a future work.\n\n\\section{summary}\\label{sec_5}\nIn summary, we suggested the transient interference in the energy domain between temporary well-separated light pulses using electronic states of band and Mott insulators as a medium, which manifests as the oscillation of the pump-probe spectrum whose frequency is indicated by Eq.~(\\ref{probe_dep}).\nThis interference could be observed only after recent developments of ultrafast spectroscopic techniques.\nThe transient interference reflects the universal property of interference between two photon pulses mediated by electron systems, which does not depend on the details of the electron systems. Therefore, the interference is also realized in the presence of electron correlation since there is a continuum structure. We examined this by calculating the pump-probe spectrum in the 1D half-filled Hubbard model.\nTo verify our prediction, we suggested an experiment for Nd$_2$CuO$_4$ with varying pump-pulse duration and delay. \nSince our theory predicts the transient oscillation even in the 1D Mott insulators, we proposed a pump-probe experiment in Sr$_2$CuO$_3$.\nFurthermore, we found that bosons coupled to electrons in the two-band model make the additional contribution to the transient interference.\nBased on the result, both magnons coupled to electrons and the continuum structure in electronic excitation spectrum would be possible origins of the oscillation observed in Nd$_2$CuO$_4$.\n\n\\begin{acknowledgments}\nWe would like to thank H. Okamoto, T. Miyamoto, I. Eremin, and P. Prelov\\v{s}ek for fruitful discussions. This work was supported by CREST (Grant No. JPMJCR1661), the Japan Science and Technology Agency, the creation of new functional devices and high-performance materials to support next-generation industries (CDMSI), and the challenge of basic science exploring extremes through multi-physics and multi-scale simulations to be tackled by using a post-K computer.\n\\end{acknowledgments}\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} {"text":"\\section{Conclusion and Future Work}\n\\label{sec:conclusions}\n\nWe presented the first distributed optimal primal-dual resource allocation\nalgorithm for uplink OFDM systems. The key features of the proposed\nalgorithm include: (a) incorporating practical OFDM system\nconstraints such as self-noise and maximum SNR constraints, (b)\nreduced primal-dual algorithm which eliminates unnecessary variable\nupdates, (c) distributed implementation by the end users and base\nstation, (d) simple local updates with limited message passing, (e)\nglobal convergence despite of the existence of multiple global\noptimal solutions, and (f) fast convergence compared with the\nstate-of-art centralized optimal algorithm. Currently we are\nextending this work in several directions, including proving the\ntheoretical convergence speed of the algorithm.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}