diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzeiqz" "b/data_all_eng_slimpj/shuffled/split2/finalzzeiqz" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzeiqz" @@ -0,0 +1,5 @@ +{"text":"\\subsection*{\\underline{#1}\\nopunct}}\n\\newcommand\\repart[1]{\\mathrm{Re}\\left[ #1 \\right]}\n\\newcommand\\impart[1]{\\mathrm{Im}\\left[ #1 \\right]}\n\\newcommand\\syl[1]{\\| #1 \\|_{\\mathrm{syl}}}\n\\newcommand\\stl[1]{\\| #1 \\|_{\\mathrm{*}}}\n\\newcommand\\ba{\\begin{align*}}\n\\newcommand\\ea{\\end{align*}}\n\\newcommand\\be{\\begin{enumerate}}\n\\newcommand\\ee{\\end{enumerate}}\n\\newcommand\\bp{\\begin{proof}}\n\\newcommand\\ep{\\end{proof}}\n\\newcommand\\bpp{\\begin{prop}}\n\\newcommand\\epp{\\end{prop}}\n\\newcommand\\bpb{\\begin{prob}}\n\\newcommand\\epb{\\end{prob}}\n\\newcommand\\bd{\\begin{defn}}\n\\newcommand\\ed{\\end{defn}}\n\\newcommand\\bh{\\begin{hint}}\n\\newcommand\\eh{\\end{hint}}\n\n\n\\newcommand\\sgn{\\mathrm{sgn}}\n\\newcommand\\stab{\\mathrm{Stab}}\n\\newcommand\\fform[1]{\\langle\\!\\langle #1\\rangle\\!\\rangle}\n\\newcommand\\vform[1]{\\vert #1\\vert}\n\\newcommand\\vstar[1]{\\| #1\\|_{\\mathrm{*}}}\n\\newcommand\\vh[1]{\\| #1\\|_{\\mathcal{H}}}\n\n\\newcommand\\bC{\\mathbb{C}}\n\\newcommand\\bE{\\mathbb{E}}\n\\newcommand\\bN{\\mathbb{N}}\n\\newcommand\\N{\\mathbb{N}}\n\\newcommand\\bR{\\mathbb{R}}\n\\newcommand\\R{\\mathbb{R}}\n\\newcommand\\bQ{\\mathbb{Q}}\n\\newcommand\\Q{\\mathbb{Q}}\n\\newcommand\\bZ{\\mathbb{Z}}\n\\newcommand\\Z{\\mathbb{Z}}\n\\newcommand\\bH{\\mathbb{H}}\n\\renewcommand\\AA{\\mathcal{A}}\n\\newcommand\\BB{\\mathcal{B}}\n\\newcommand\\CC{\\mathcal{C}}\n\\newcommand\\FF{\\mathcal{F}}\n\\newcommand\\GG{\\mathcal{G}}\n\\newcommand\\DD{\\mathcal{D}}\n\\newcommand\\HH{\\mathcal{H}}\n\\newcommand\\KK{\\mathcal{K}}\n\\newcommand\\XX{\\mathcal{X}}\n\\newcommand\\sech{\\operatorname{sech}}\n\\newcommand\\Sym{\\operatorname{Sym}}\n\n\\newcommand\\ev{\\operatorname{ev}}\n\\newcommand\\cay{\\operatorname{Cayley}}\n\\newcommand\\aut{\\operatorname{Aut}}\n\\newcommand\\Inn{\\operatorname{Inn}}\n\\newcommand\\Mono{\\operatorname{Mono}}\n\\newcommand\\PMod{\\operatorname{PMod}}\n\\newcommand\\Hom{\\operatorname{Hom}}\n\\DeclareMathOperator\\Push{\\mathcal{Push}}\n\\DeclareMathOperator\\Forget{\\mathcal{Forget}}\n\\newcommand\\Isom{\\operatorname{Isom}}\n\\newcommand\\UT{\\operatorname{UT}}\n\\newcommand\\Comm{\\operatorname{Comm}}\n\\newcommand\\Symp{\\operatorname{Symp}}\n\\newcommand\\Fill{\\operatorname{Fill}}\n\\newcommand\\supp{\\operatorname{supp}}\n\\newcommand\\Id{\\operatorname{Id}}\n\\newcommand\\lk{\\operatorname{Lk}}\n\\newcommand\\st{\\operatorname{St}}\n\\newcommand\\SL{\\operatorname{SL}}\n\\newcommand\\Diffb{\\operatorname{Diff}_+^{1+\\mathrm{bv}}}\n\\newcommand\\Cb{C^{1+\\mathrm{bv}}}\n\\newcommand\\diam{\\operatorname{diam}}\n\\newcommand\\Gam{\\Gamma}\n\\newcommand\\gam{\\Gamma}\n\\newcommand\\Mod{\\operatorname{Mod}}\n\\newcommand\\PSL{\\operatorname{PSL}}\n\\newcommand\\symp{\\operatorname{Symp}}\n\\DeclareMathOperator\\Homeo{Homeo}\n\\newcommand\\mC{\\mathcal{C}}\n\\newcommand\\mL{\\mathcal{L}}\n\\newcommand\\mM{\\mathcal{M}}\n\\newcommand\\xek{(X^e)_k}\n\\newcommand\\yt{\\widetilde}\n\\newcommand\\sse{\\subseteq}\n\\newcommand\\co{\\colon}\n\\DeclareMathOperator\\tr{tr}\n\\DeclareMathOperator\\Fix{Fix}\n\\DeclareMathOperator\\Out{Out}\n\\DeclareMathOperator\\Aut{Aut}\n\\DeclareMathOperator\\CAT{CAT}\n\\DeclareMathOperator\\Diff{Diff}\n\\DeclareMathOperator\\Exp{Exp}\n\\newcommand\\catz{\\mathrm{CAT}(0)}\n\\DeclareMathOperator\\Stab{Stab}\n\\DeclareMathOperator\\Orbit{Orbit}\n\\newcommand\\var{\\operatorname{var}}\n\\newcommand\\rot{\\operatorname{rot}}\n\\newcommand\\Cbv{C^{1+\\mathrm{bv}}}\n\\DeclareMathOperator\\Per{Per}\n\n\n\n\\renewcommand{\\thefootnote}{\\alpha{footnote}}\n\\newcommand{\\arxiv}[1]\n{\\texttt{\\href{http:\/\/arxiv.org\/abs\/#1}{arXiv:#1}}}\n\\newcommand{\\doi}[1]\n{\\texttt{\\href{http:\/\/dx.doi.org\/#1}{doi:#1}}}\n\\renewcommand{\\MR}[1]\n{\\href{http:\/\/www.ams.org\/mathscinet-getitem?mr=#1}{MR#1}}\n\n\n\n\n\n\n\\def{Free products and the algebraic structure of diffeomorphism groups}{{Free products and the algebraic structure of diffeomorphism groups}}\n\\def{Sang-hyun Kim and Thomas Koberda}{{Sang-hyun Kim and Thomas Koberda}}\n\\usepackage{hyperref}\n\\hypersetup{\n \n colorlinks=false,\n plainpages,\n urlcolor=black,\n linkcolor=black\n pdftitle= {Free products and the algebraic structure of diffeomorphism groups},\n pdfauthor= {{Sang-hyun Kim and Thomas Koberda}}\n}\n\n\n\n\\theoremstyle{theorem}\n\\newtheorem{thm}{Theorem}[section]\n\\newtheorem{lem}[thm]{Lemma}\n\\newtheorem{cor}[thm]{Corollary}\n\\newtheorem{prop}[thm]{Proposition}\n\\newtheorem{con}[thm]{Conjecture}\n\\newtheorem{que}[thm]{Question}\n\\newtheorem*{claim*}{Claim}\n\\newtheorem*{fact}{Fact}\n\\newtheorem{claim}{Claim}\n\\newtheorem{thmA}{Theorem}\n\\renewcommand*{\\thethmA}{\\Alph{thmA}}\n\\newtheorem{corA}[thmA]{Corollary}\n\\renewcommand*{\\thecorA}{\\Alph{corA}}\n\n\n\\theoremstyle{remark}\n\\newtheorem{exmp}[thm]{Example}\n\\newtheorem{rem}[thm]{Remark}\n\n\n\\theoremstyle{definition}\n\\newtheorem{defn}[thm]{Definition}\n\\newtheorem{prob}{Problem}[section]\n\\newtheorem{exc}{Exercise}[section]\n\n\\begin{document}\n\\title{Free products and the algebraic structure of diffeomorphism groups}\n\\date{\\today}\n\\keywords{free product; metabelian group; Thompson's group; right-angled Artin group; smoothing; co-graph}\n\\subjclass[2010]{Primary: 57M60; Secondary: 20F36, 37C05, 37C85, 57S05}\n\n\\author[S. Kim]{Sang-hyun Kim}\n\\address{Department of Mathematical Sciences, Seoul National University, Seoul, Korea}\n\\email{s.kim@snu.ac.kr}\n\\urladdr{http:\/\/cayley.kr}\n\n\\author[T. Koberda]{Thomas Koberda}\n\\address{Department of Mathematics, University of Virginia, Charlottesville, VA 22904-4137, USA}\n\\email{thomas.koberda@gmail.com}\n\\urladdr{http:\/\/faculty.virginia.edu\/Koberda}\n\n\n\n\n\n\\begin{abstract}\nLet $M$ be a compact one--manifold,\nand let $\\Diffb(M)$ denote the group of $C^1$ orientation preserving diffeomorphisms of $M$ whose first derivatives have bounded variation.\nWe prove that if $G$ is a group which is not virtually metabelian, then $(G\\times\\Z)*\\Z$ is not realized as a subgroup of $\\Diffb(M)$.\nThis gives the first examples of finitely generated groups $G,H\\le \\Diff_+^\\infty(M)$ such that $G\\ast H$ does not embed into $\\Diffb(M)$.\nBy contrast, \nfor all countable groups $G,H\\le\\Homeo^+(M)$ there exists an embedding $G\\ast H\\to \\Homeo^+(M)$.\nWe deduce that many common groups of homeomorphisms do not embed into $\\Diffb(M)$, for example the free product of $\\bZ$ with Thompson's group $F$.\nWe also complete the classification of right-angled Artin groups which can act smoothly on $M$\nand in particular, recover the main result of a joint work of the authors with Baik~\\cite{BKK2014}. \nNamely, a right-angled Artin group $A(\\gam)$ either admits a faithful $C^{\\infty}$ action on $M$, or $A(\\gam)$ admits no faithful $\\Cb$ action on $M$. In the former case, $A(\\gam)\\cong\\prod_i G_i$\nwhere $G_i$ is a free product of free abelian groups.\nFinally, we develop a hierarchy of right-angled Artin groups, with the levels of the hierarchy corresponding to the number of semi-conjugacy classes of possible actions of these groups on $S^1$.\n\\end{abstract}\n\n\\maketitle\n\n\n\n\\section{Introduction}\n\nLet $M$ be a compact one--manifold. In this article, we study the algebraic structure of the group $\\Diffb(M)$, where here $\\Diffb(M)$ denotes the group of $C^1$ diffeomorphisms of $M$ whose derivatives have bounded variation.\nSpecifically, we consider the restrictions placed on subgroups of $\\Diffb(M)$ by the $\\Cb$ regularity assumption. Our main result implies that there is a large class $\\AA_0 $ of finitely generated subgroups of $\\Diff_+^\\infty(M)$ such that for all $G,H\\in\\AA_0$, the free product $G\\ast H$ can never be realized as a subgroup of $\\Diffb(M)$; see Corollary~\\ref{cor:classification}.\n\nAs a corollary, we complete a program initiated by Baik and the authors in~\\cite{BKK2014,BKK2016} to decide which right-angled Artin groups admit faithful $C^{\\infty}$ actions on a compact one--manifold, and exhibit many classes of finitely generated subgroups of $\\Homeo^+(M)$ which cannot be realized as subgroups of $\\Diffb(M)$.\n\n\\subsection{Statement of results}\n\nUnless otherwise noted, $M$ will denote a compact one--manifold.\n That is to say, $M$ is a finite union of disjoint closed intervals $I=[0,1]$ and circles $S^1=\\bR\/\\bZ$.\nAn \\emph{action} on a one--manifold is always assumed to mean \nan orientation preserving action. \nFor each $0\\le r\\le\\infty$ or $r=\\omega$, we let $\\Diff^r_+(M)$ denote the group of orientation preserving $C^r$ (analytic if $r=\\omega$) diffeomorphisms.\nWe write $\\Homeo^+(M)=\\Diff^0_+(M)$.\nOur main result is the following:\n\n\\begin{thm}\\label{thm:main}\nIf $G$ is a group which is not virtually metabelian, then the group $(G\\times\\Z)*\\Z$ admits no faithful $C^{1+\\mathrm{bv}}$ action on $M$.\n\\end{thm}\n\n\nTheorem~\\ref{thm:main} clearly implies that $(G\\times\\Z)*\\Z$ does not embed into $\\Diff^r_+(M)$ for every level of regularity $r\\geq 2$.\n A key step in our proof of Theorem~\\ref{thm:main} is the following result on $C^1$--smoothability:\n\\begin{thm}[Theorem~\\ref{t:tech-main}]\\label{thm:tech-main-intro}\nLet $X\\in\\{I,S^1\\}$,\nand let\n $a,b,t\\in\\Diff^1_+(X)$.\n If\n\\[\\supp a\\cap\\supp b=\\varnothing,\\] \nthen the group $\\form{a,b,t}$ is not isomorphic to $\\bZ^2\\ast\\bZ$.\n\\end{thm}\n\n\nWe have stated Theorem~\\ref{thm:main} as above for clarity and concision, though several stronger statements can be deduced from the proof we give. In the particular case of finitely generated groups, we note the following, which should be compared to Corollary~\\ref{cor:cnt}:\n\n\n\\begin{cor}\\label{cor:fg1}\nIf $G$ is a finitely generated group which is not virtually abelian, then $(G\\times\\Z)*\\Z$ admits no faithful $\\Cb$ action on $M$. \n\\end{cor}\nThe motivation for proving Theorem~\\ref{thm:main} came from investigating right-angled Artin subgroups of $\\Diffb(M)$ (cf. Corollary~\\ref{cor:classification} below). The simplest right-angled Artin group to which Theorem~\\ref{thm:main} applies is $(F_2\\times\\Z)*\\Z$:\n\n\n\\begin{cor}\\label{cor:FtimesZ}\nThe group $(F_2\\times\\Z)*\\Z$ is not a subgroup of $\\Diffb(M)$.\n\\end{cor}\n\nSince the hypotheses on Theorem~\\ref{thm:main} are relatively weak, there are many finitely generated subgroups of $\\Homeo^+(M)$ which can be thus shown to admit no faithful $\\Cb$ actions on a compact manifold. Recall that \\emph{Thompson's group $F$} is a group of piecewise linear homeomorphisms of the interval, with dyadic break points and with all slopes being powers of two. \\emph{Thompson's group $T$} is the analogous group defined for the circle. A famous result of Ghys--Sergiescu~\\cite{GS1987} says that the standard actions of $T$ and $F$ are topologically conjugate into a group of $C^{\\infty}$ diffeomorphisms of the circle and of the interval, respectively. However, we have the following:\n\n\\begin{cor}\\label{cor:Thompson}\nThe groups $F*\\Z$ and $T*\\Z$ are not subgroups of $\\Diffb(M)$.\n\\end{cor}\n\nIn the vein of Corollary~\\ref{cor:Thompson}, we do not know the answer to the following:\n\n\\begin{que}\\label{que:thompson-critical}\nAre the groups $F*\\Z$ and $T*\\Z$ subgroups of $\\Diff_+^1(I)$ and $\\Diff_+^1(S^1)$ respectively? If so, what is their optimal regularity, i.e. the supremum of the H\\\"older continuity exponent $\\tau\\in [0,1)$ such that these groups embed in $\\Diff_+^{1+\\tau}$ for the relevant manifolds (cf.~\\cite{JNR2018,KK2017crit})?\n\\end{que}\n\n\n\n\nReturning to the original motivation, Theorem \\ref{thm:main} combined with results of Farb--Franks and Jorquera completes the classification of right-angled Artin groups admitting actions of various regularities on compact one--manifolds. Before stating the result, we define some terminology. We write $\\gam$ for a finite simplicial graph with vertex set $V(\\gam)$ and edge set $E(\\gam)$. The \\emph{right-angled Artin group} (or \\emph{RAAG}, for short) on $\\gam$ is defined as \\[A(\\gam)=\\langle V(\\gam)\\mid [v,w]=1 \\textrm{ if and only if }\\{v,w\\}\\in E(\\gam)\\rangle.\\]\n\nA subgraph $\\Lambda$ of a graph $\\gam$ is called a \\emph{full} subgraph if $\\Lambda$ is spanned by the vertices of $\\Lambda$. That is, two vertices in $\\Lambda$ are adjacent if and only if they are adjacent in $\\gam$.\nA simplicial graph $\\gam$ is called \\emph{$P_4$--free} if no full subgraph of $\\gam$ is isomorphic to a path $P_4$ on four vertices. Such graphs are often called \\emph{cographs}~\\cite{CLB1981,KK2013}. We write $\\mathcal{K}$ for the class of cographs. It is well--known that cographs can be fit into a hierarchy which is defined as follows:\n\\begin{enumerate}\n\\item\nThe class $\\mathcal{K}_0$ consists of a single vertex;\n\\item\nIf $n\\geq 1$ is odd then $\\mathcal{K}_n$ is obtained by taking $\\mathcal{K}_{n-1}$ together with finite joins of elements in $\\mathcal{K}_{n-1}$;\n\\item\nIf $n\\geq 2$ is even then $\\mathcal{K}_n$ is obtained by taking $\\mathcal{K}_{n-1}$ together with finite disjoint unions of elements in $\\mathcal{K}_{n-1}$.\n\\end{enumerate}\nHere, a \\emph{join} of two simplicial graphs $X$ and $Y$ is a simplicial graph consisting of the disjoint union of $X$ and $Y$, together with an edge of the form $\\{x,y\\}$ for every vertex $x$ of $X$ and every vertex $y$ of $Y$.\n\n\nWe have that $\\mathcal{K}_0\\subset\\mathcal{K}_1\\subset\\KK_2\\subset\\cdots$ and \\[\\mathcal{K}=\\bigcup_i\\mathcal{K}_i.\\] \nNote that join and disjoint union correspond to direct product and free product respectively, so that right-angled Artin groups on cographs are exactly the smallest class of groups containing $\\Z$, which is closed under finite direct products, and which is closed under finite free products. The reader will observe that if $\\gam\\in\\mathcal{K}_0$ then $A(\\gam)\\cong\\Z$. Similarly, if $\\gam\\in\\mathcal{K}_1$ then $A(\\gam)$ is free abelian, and if $\\gam\\in\\mathcal{K}_2$ then $A(\\gam)$ is a free product of free abelian groups.\nIf $\\gam\\in\\mathcal{K}_3$ then $A(\\Gamma)$ can be written as \\[A(\\Gamma)=\\prod_{i=1}^m G_i,\\]\n where each $G_i$ is a free product of free abelian groups.\nIn~\\cite{BKK2016}, Baik and the authors proved that if $A(\\gam)$ admits an injective homomorphism into $\\Diffb(M)$, then $\\gam\\in\\KK$. We will deduce a strengthening of this result, using Theorem~\\ref{thm:main}.\n\n\n\n\n\n\\begin{cor}[cf. \\cite{BKK2016}]\\label{cor:classification}\nLet $A(\\gam)$ be a right-angled Artin group.\n\\begin{enumerate}\n\\item (see \\cite{FF2003,Jorquera})\nThere exists an injective homomorphism $A(\\gam)\\to\\Diff_+^1(M)$; thus, $A(\\gam)$ admits a faithful $C^1$ action of $M$.\n\\item\nIf there exists an injective homomorphism $A(\\gam)\\to\\Diffb(M)$ then $\\gam\\in\\mathcal{K}_3$; conversely, if $\\gam\\in\\mathcal{K}_3$ then $A(\\gam)\\le\\Diff_+^{\\infty}(M)$.\n\\end{enumerate}\n\\end{cor}\n\nIn particular, if we define\n\\[\n\\AA_0=\\{A(\\Gamma)\\mid \\Gamma\\in\\KK_3\\setminus\\KK_2\\}\\]\nthen each $G\\in\\AA_0$ embeds into $\\Diff_+^\\infty(M)$,\nbut for all $G,H\\in\\AA_0$ the group $G\\ast H$ never embeds into $\\Diff_+^{1+\\mathrm{bv}}(M)$.\nBy contrast, we have the following, which is well--known from several contexts.\n\n\\begin{prop}[cf. \\cite{KM1996,Rivas2012JAlg,BS2015}]\\label{prop:free prod}\nThe class of countable subgroups of $\\Homeo^+(M)$ is closed under finite free products.\n\\end{prop}\n\nIn the same spirit of Question~\\ref{que:thompson-critical} and the authors' paper~\\cite{KK2017crit}, we have the following:\n\n\\begin{que}\nLet $\\gam\\notin\\mathcal{K}_3$. What is the supremum of $\\tau\\in [0,1)$ for which $A(\\gam)$ embeds in $\\Diff_+^{1+\\tau}(M)$? Does $\\tau$ depend on $\\gam$?\n\\end{que}\n\nLet $S$ be an orientable surface of genus $g$ and with $n$ punctures or boundary components. We say that $S$ is \\emph{sporadic} if \\[3g-3+n\\leq 1.\\] We write $\\Mod(S)$ for the mapping class group of $S$, i.e. $\\Mod(S)=\\pi_0(\\Homeo^+(S))$. Using the main result of~\\cite{Koberda2012}, we immediately recover the following result as a corollary of Theorem~\\ref{thm:main}:\n\n\\begin{cor}[cf.~\\cite{BKK2016}]\nLet $M$ be a compact one--manifold, and let $S$ be an orientable finite-type surface. Then there exists a finite index subgroup $G\\le\\Mod(S)$ such that $G\\le\\Diffb(M)$ if and only if $S$ is sporadic.\n\\end{cor}\n\n\n\nTheorem \\ref{thm:main} allows us to build a hierarchy on right-angled Artin groups, whose levels correspond to right-angled Artin groups with more or fewer ``dynamically different\" actions on the circle. Roughly speaking, two group actions \\[\\rho_1,\\rho_2\\colon G\\to\\Homeo^+(S^1)\\] are \\emph{semi--conjugate} (or, \\emph{monotone-equivalent}) if there exists \nanother action \\[\\rho\\co G\\to\\Homeo^+(S^1)\\]\nand monotone degree one maps $h_i\\colon S^1\\to S^1$ such that\n\\[h_i\\circ \\rho=\\rho_i\\circ h_i\\] for each $i=1,2$. See~\\cite{Ghys1987,CD2003IM,Ghys2001,Mann-hb,BFH2014,KKMj2016} for instance, and the many references therein.\nA \\emph{projective action} of a group $G$ is a representation\n\\[\n\\rho\\co G\\to \\PSL_2(\\bR),\\]\nwhere $\\PSL_2(\\R)$ sits inside of $\\Homeo^+(S^1)$ as the group of projective analytic diffeomorphisms of $S^1$. \n\n\n\\begin{cor}\\label{cor:conj}\nLet $A(\\gam)$ be a right-angled Artin group.\n\\begin{enumerate}\n\\item\nIf $\\gam\\in\\mathcal{K}_2$, then $A(\\gam)$ admits uncountably many distinct semi--conjugacy classes of faithful orientation preserving projective actions on $S^1$;\n\\item\nIf $\\gam\\in\\mathcal{K}_3\\setminus\\mathcal{K}_2$ then any faithful orientation preserving $\\Cb$ action of $A(\\gam)$ on $S^1$ has a periodic point and no dense orbits and hence admits at most countably many distinct semi--conjugacy classes of $\\Cb$ actions on $S^1$;\n\\item\nIf $\\gam\\notin\\mathcal{K}_3$ then $A(\\gam)$ admits no faithful $\\Cb$ action on $S^1$.\n\\end{enumerate}\n\\end{cor}\n\nIn the case of analytic actions on a compact connected one--manifold $M$, one has the following result of Akhmedov and Cohen:\n\n\\begin{thm}[See~\\cite{AC2015TA}]\nThe right-angled Artin group $A(\\gam)$ embeds into $\\Diff^\\omega(M)$ if and only if $\\Gamma\\in\\KK_2$, i.e. $A(\\gam)$ decomposes as a free product of free abelian groups.\n\\end{thm}\n\n\\subsection{Notes and references}\n\nThis paper\nreveals some of the subtlety of the interplay between algebra and regularity in diffeomorphism groups of one--manifolds. \nOur paper arose during the effort to complete the classification of right-angled Artin subgroups of $\\Diff_+^{\\infty}(S^1)$\nin the spirit of~\\cite{BKK2016}.\nThe essential content of this paper is Lemma~\\ref{l:main}, which exhibits an explicit element of the kernel of any given $\\Cb$ action of $(G\\times\\Z)*\\Z$ action on a compact one--manifold.\nSince a group of the form $(G\\times\\Z)*\\Z$ is generally simpler than the right-angled Artin groups considered in~\\cite{BKK2016}, it is more difficult to find elements in the kernel of a given action, and therefore we develop more sophisticated tools here. We note that our main result does subsume the main result of~\\cite{BKK2016}. Indeed, the main result of~\\cite{BKK2016} is that there is no injective homomorphism $A(P_4)\\to\\Diffb(M)$, where here \\[A(P_4)=\\langle a,b,c,d\\mid [a,b]=[b,c]=[c,d]=1\\rangle.\\] The group $A(P_4)$ contains a copy of $(F_2\\times\\Z)*\\Z$, which cannot embed in $\\Diffb(M)$ by Corollary~\\ref{cor:FtimesZ}. An explicit embedding of $(F_2\\times\\Z)*\\Z$ into $A(P_4)$ is given by $\\langle a,b,c,dad^{-1}\\rangle\\le A(P_4)$ (see~\\cite{KK2013} for a discussion of this fact).\n\nThe program completed by Corollary~\\ref{cor:classification} fully answers a question raised in a paper of M. Kapovich (attributed to Kharlamov) as to which right-angled Artin groups admit faithful $C^{\\infty}$ actions on the circle~\\cite{Kapovich2012}.\n\nRight-angled Artin subgroups of diffeomorphism groups of one--manifolds find an analogue in right-angled Artin subgroups of linear groups. It is well--known that right-angled Artin groups are always linear over $\\Z$ and hence admit injective homomorphisms into $\\SL_n(\\Z)$~\\cite{Humphries1994,HW1999,DJ2000}. For $\\SL_3(\\Z)$, it is still unclear which right-angled Artin groups appear as subgroups. Long--Reid~\\cite{LR2011} showed that $F_2\\times\\Z$ is not a subgroup of $\\SL_3(\\Z)$, which implies that any right-angled Artin subgroup of $\\SL_3(\\Z)$ is a free product of free abelian groups of rank at most two. It is currently unknown whether or not $\\Z^2*\\Z$ is a subgroup of $\\SL_3(\\Z)$.\n\nA consequence of the technical work behind Theorem~\\ref{thm:main} is a certain criterion to prove that a group contains a lamplighter subgroup. See Lemma~\\ref{l:recursive} and Proposition~\\ref{t:tech-main2} for precise statements.\n\nA crucial step in our proof of the main theorem is a $C^1$--rigidity result, which is Theorem~\\ref{t:tech-main}. \nWe note that Thurston~\\cite{Thurston1974Top}, Calegari~\\cite{Calegari2008AGT}, Navas~\\cite{Navas2008GAFA} and Bonatti, Monteverde, Navas and Rivas~\\cite{BMNR2017MZ} explored various remarkable $C^1$--rigidity results. \nIn particular, a $C^1$--rigidity result on Baumslag--Solitar groups in~\\cite{BMNR2017MZ} was employed in a very recent paper by Bonatti, Lodha and Triestino~\\cite{BLT2017},\nto produce certain piecewise affine homeomorphism groups of $\\bR$ which do not embed into $\\Diff^1_+(I)$.\n\nAs for other classes of groups of homeomorphisms which cannot be realized as groups of $\\Cb$ diffeomorphisms, Corollary~\\ref{cor:Thompson} is a complement to a result of the second author with Lodha~\\cite{KL2017}, in which they show that certain ``square roots\" of Thompson's group $F$ may fail to act faithfully by $\\Cb$ diffeomorphisms on a compact one--manifold, even though they are manifestly groups of homeomorphisms of these manifolds. Thus, the Ghys--Sergiescu Theorem appears to place $F$ and $T$ at the cusp of smoothability in the sense that even relatively minor algebraic variations on $F$ and on $T$ fail to be smoothable.\n\nFinally, we remark on the optimality of the differentiability hypothesis. \nCorollary~\\ref{cor:classification} shows that every right-angled Artin group can act faithfully by $C^1$ diffeomorphisms, but nearly none of them can act by $\\Cb$ diffeomorphisms. \nAs for Corollary~\\ref{cor:not closed} and Proposition~\\ref{prop:free prod}, we have the following result (based on~\\cite{BMNR2017MZ} and on a suggestion by Navas), which is proved in the authors' recent manuscript~\\cite{KK2017crit}:\n\n\\begin{prop}[\\cite{KK2017crit}]\\label{prop:c1freeproduct}\nLet $M\\in\\{I,S^1\\}$, and let $BS(1,2)$ be the Baumslag--Solitar group with the presentation $\\langle s,t\\mid sts^{-1}=t^2\\rangle$. Then the group $(BS(1,2)\\times\\Z)*\\Z$ is not a subgroup of $\\Diff_+^1(M)$. In particular, the class of finitely generated subgroups of $\\Diff_+^1(M)$ is not closed under taking finite free products.\n\\end{prop}\n\n\\section{Background on one--dimensional smooth dynamics}\nWe very briefly summarize the necessary background from one--dimensional dynamics. The reader may also consult~\\cite{BKK2016}, parts of which we repeat here, and where we direct the reader for proofs.\n\n\\subsection{Poincar\\'e's theory of rotation numbers}\n\nLet $f\\in\\Homeo^+(S^1)$, and let $\\tilde f\\co \\bR\\to\\bR$ be an arbitrary lift of $f$.\nThen the \\emph{rotation number} of $f$ is defined as\n\\[\n\\rot f = \\lim_{n\\to\\infty}\\frac{\\tilde f^n(x)}{n}\\in \\bR\/\\bZ=S^1\\]\nwhere $x\\in \\bR$. Then $\\rot f$ is well-defined, and independent of the choice of a lift $\\tilde f$ and a base point $x\\in \\bR$; see~\\cite{Navas2011} for instance. \nThe set of periodic points of $f$ is denoted as $\\Per f$.\nLet us record some elementary facts.\n\\begin{lem}\\label{l:inv}\nFor $f\\in\\Homeo^+(S^1)$, the following hold.\n\\be\n\\item\n$\\rot(f) =0$ if and only if $\\Fix f\\ne\\varnothing$.\n\\item\n$\\rot(f)\\in\\bQ$ if and only if $\\Per f\\ne\\varnothing$.\n\\item\\label{p:inv}\nIf $x\\in S^1$ and $g\\in\\Homeo^+(S^1)$ satisfy\n\\[\nf^n(x)=g^n(x)\\]\nfor all $n\\in\\bZ$, then $\\rot(f)=\\rot(g)$.\n\\ee\n\\end{lem}\n\nThe rotation number is a continuous class function (that is, constant on each conjugacy class)\n\\[\n\\rot\\co \\Homeo^+(S^1)\\to S^1.\\]\nMoreover, the rotation number restricts to a group homomorphism on each amenable subgroup of $\\Homeo^+(S^1)$; see~\\cite{Ghys2001}.\n\nLet us recall the following classical result.\n\n\n\\begin{thm}[H\\\"older's Theorem~\\cite{Holder1901}; see~\\cite{Navas2011}]\\label{t:hoelder}\nA group acting freely on $\\bR$ or on $S^1$ by orientation preserving homeomorphisms is abelian.\n\\end{thm}\n\nWe will use the following variation, which is similar to \\cite[Theorem 2.2]{FF2001}.\n\\begin{cor}[cf.~\\cite{FF2001}]\\label{c:hoelder-s1}\nLet $X$ be a nonempty closed subset of $S^1$,\nand let $G$ be a group acting freely on $X$ by orientation preserving homeomorphisms.\nThen the action of $G$ extends to a free action $\\rho\\co G\\to\\Homeo^+(S^1)$ such that \n\\[\\rot\\circ\\rho\\co G\\to S^1\\] is an injective group homomorphism.\n\\end{cor}\n\\bp\nIf $(a,b)$ is a component of $S^1\\setminus X$ and $g \\in G$, then $(ga,gb)$ is also a component of $S^1\\setminus X$. So, $G$ extends to some action $\\rho$ on $S^1$ in an affine manner. \nPut \\[S^1\\setminus X = \\coprod_{i\\ge1} I_i.\\] \nIf $\\rho (g)y = y$ for some $g \\in G$ and for some $y \\in S^1$, then $y \\in I_i$ for some $i$. We have that $ \\rho (g)$ restricts to the identity on $ I_i$ by definition. This would imply $\\rho (g)\\partial I_i = g \\partial I_i = \\partial I_i$, and so, $g = 1$. That is, the action $\\rho$ of $G$ on $S^1$ is free.\n\nBy H\\\"older's theorem, we see $\\rho(G)\\cong G$ is abelian. Since abelian groups are amenable, \nwe have that $\\rot \\circ \\rho$ is a group homomorphism. The freeness of the action $\\rho$ implies that $\\rot\\circ\\rho(g)\\ne0$ for all nontrivial $g$.\\ep\n\n\n\n\\subsection{Kopell--Denjoy theory}\nLet $M\\in\\{I,S^1\\}$.\nWe denote by $\\var(g;M)$ the total variation of a map $g\\co M\\to\\bR$:\n\\[\\var(g;M) = \\sup\\left\\{\\sum_{i=0}^{n-1}\\left| g(a_{i+1})- g(a_i)\\right|\n\\co (a_i\\co 0\\le i\\le n)\\text{ is a partition of }M\\right\\}.\\]\nIn the case $M=S^1$, we require $a_n=a_0$ in the above definition.\nFollowing \\cite{Navas2011},\nwe say a $C^1$ diffeomorphism $f$ on $M$ is $\\Cbv$ if $\\var(f';M)<\\infty$.\nWe let $\\Diffb(M)$ denote\nthe group of orientation preserving $\\Cbv$ diffeomorphisms of $M$.\nThe following two results play a fundamental role on the study of $\\Cbv$ diffeomorphisms.\n\n\\begin{thm}[Denjoy's Theorem {\\cite{Denjoy1958}, \\cite{Navas2011}}]\\label{t:denjoy}\nIf $a\\in\\Diffb(S^1)$ and $\\Per a=\\emptyset$, then $a$ is topologically conjugate to an irrational rotation.\n\\end{thm}\n\n\\begin{thm}[Kopell's Lemma {\\cite{Kopell1970}, \\cite{Navas2011}}]\\label{t:kopell}\nSuppose $a\\in\\Diffb[0,1)$, $b\\in\\Diff_+^1[0,1)$, and $[a,b]=1$.\nIf $\\Fix a\\cap (0,1)=\\varnothing$ and $b\\ne 1$,\nthen $\\Fix b\\cap (0,1)=\\varnothing$.\n\\end{thm}\nWe remark that the original statement by Kopell was for $C^2$--regularity. Navas extended her result to $C^{1+\\mathrm{bv}}$ case~\\cite[Theorem 4.1.1]{Navas2011}.\n\nLet $X$ be a topological space. Then we define the \\emph{support} of $h\\in \\Homeo(X)$ as\n\\[\\supp h = X\\setminus\\Fix h.\\] \nIt is convenient for us to consider the non--standard \\emph{open support} of a homeomorphism as defined here, whereas many other authors use the closure of the open support. We will consistently mean the open support unless otherwise noted.\n\nFor a subgroup $G\\le\\Homeo(X)$, we put\n\\[\\supp G = \\bigcup_{g\\in G}\\supp g.\\]\nWe say that $f\\in\\Homeo(X)$ is \\emph{grounded} if $\\Fix f\\neq\\varnothing$. We note that every $f\\in\\Homeo^+(I)$ is grounded by definition.\n\nThe following important observations on commuting $\\Cbv$ diffeomorphisms essentially builds on Kopell's Lemma and H\\\"older's Theorem. \n\n\n\\begin{lem}\\label{l:disj-abel}\nThe following hold:\n\\be\n\\item (Disjointness Condition,~\\cite{BKK2016})\nLet $M\\in\\{I,S^1\\}$,\nand let $a,b\\in\\Diffb(M)$ be commuting grounded diffeomorphisms.\nIf $A$ and $B$ are components of $\\supp a$ and $\\supp b$ respectively,\nthen either $A=B$ or $A\\cap B=\\varnothing$.\n\\item (Abelian Criterion, cf.~\\cite{FF2001})\nIf $a,b,c\\in\\Diffb(I)$ satisfy $\\Fix a =\\partial I$ and that\n$[a,b]=1=[a,c]$, then $[b,c]=1$.\n\\ee\n\\end{lem}\n\n\\begin{rem}\nThe abelian criterion as given by Farb and Franks is a straightforward consequence of H\\\"older's Theorem and Kopell's Lemma. We thank one of the referees for pointing out another simple proof using Szekeres' Theorem~\\cite{Navas2011}.\n\\end{rem}\n\nThe notation $f\\restriction_A$ for a function $f$ (or set of functions) and a set $A$ means the restriction to $A$.\nWe will need the following properties of centralizer groups.\n\n\n\\begin{lem}\\label{l:circle}\nLet $a\\in \\Diffb(S^1)$ be an infinite order element,\nand let $Z(a)$ be the centralizer of $a$ in $\\Diffb(S^1)$.\nThen the following hold.\n\\be\n\\item\\label{p:zz}\nIf $a$ is grounded and if a group $H\\le Z(a)$ is generated by grounded elements, then every element in $H$ is grounded and moreover,\n\\[\\supp a\\cap \\supp [H,H]=\\varnothing.\\]\n\\item\\label{p:rot-irr}\nIf $\\rot a\\not\\in\\bQ$, then $Z(a)$ is topologically conjugate to a subgroup of $\\operatorname{SO}(2,\\bR)$.\n\\item\\label{p:rot-q}\nIf $\\rot a\\in\\bQ$, then $\\rot Z(a)\\sse\\bQ$.\n\\item\\label{p:hom} \n({cf. \\cite[Lemma 3.4]{FF2001}})\nThe rotation number restricts to a homomorphism on $Z(a)$;\nin particular, every element of $[Z(a),Z(a)]$ is grounded.\n\\ee\n\\end{lem}\n\\bp\n(\\ref{p:zz})\nLet $J$ be a component of $\\supp a$.\nBy the Disjointness Condition, the group $H$ acts on the open interval $J$. Since $H$ fixes $\\partial J$, every element of $H$ is grounded.\nThe Abelian Criterion implies that\n\\[[H,H]\\restriction_J=1.\\]\nSo, we have that $J\\cap \\supp[H,H]=\\varnothing$.\n\n(\\ref{p:rot-irr}) By Denjoy's Theorem, the map $a$ is topologically conjugate to an irrational rotation. The centralizer of an irrational rotation in $\\Homeo^+(S^1)$ is $\\operatorname{SO}(2,\\bR)$;\nsee~\\cite[Proposition 2.10]{FF2001} or~\\cite[Exercise 2.2.12]{Navas2011}. \n\n\n\n(\\ref{p:rot-q})\nSuppose some element $b$ in $Z(a)$ has an irrational rotation number. \nSince $a\\in Z(b)$, part (\\ref{p:rot-irr}) implies that $a$ is conjugate to a rotation. This is a contradiction, for a rotation with a rational rotation number must have a finite order.\n\n(\\ref{p:hom})\nIf $\\rot a$ is irrational, then the conclusion follows from part (\\ref{p:rot-irr}). So we may assume $\\rot a\\in\\bQ$. Then $a^p$ is grounded for some $p\\ne0$.\nSince $Z(a)\\le Z(a^p)$, it suffices to prove the lemma for $a^p$. In other words, we may further suppose that $a$ is grounded.\n\nLet us put \n\\[G = Z(a),\\quad G_0 = \\rot^{-1}(0)\\cap G.\\]\nPart (\\ref{p:zz}) implies that \n\\[\nG_0=\\bigcup_{x\\in S^1}\\stab_G(x)=\\form{G_0}\\]\nis a group.\nSince $\\rot$ is a class function on $\\Homeo^+(S^1)$,\nwe see that $G_0\\unlhd G$.\n\nFor each $x\\in \\Fix a$ and $g\\in G$, we note\n\\[\nag(x) = ga(x) = g(x).\\]\nSo, $\\Fix a$ is $G$--invariant. \nWe have a nonempty proper closed $G$--invariant set\n\\[\nX=\\partial \\Fix a.\\]\n\\begin{claim}\\label{claim:HG}\nFor all $x\\in X$, the group $G_0$ fixes $x$.\n\\end{claim}\nFor each $J\\in\\pi_0\\supp a$ and $g\\in G_0$, \nwe have seen in part (\\ref{p:zz}) that $\\partial J\\sse \\Fix g$.\nSince we can write\n\\[\nX = \\overline{\\bigcup\\{\\partial J\\mid J\\in\\pi_0\\supp a\\}}\\]\nwe see that $X\\sse \\Fix g$. This proves the claim.\n\nLet $p\\co G\\to G\/G_0$ denote the quotient map.\nBy Claim~\\ref{claim:HG}, \n the natural action\n\\[G\/G_0\\to \\Homeo^+(X),\\quad p(g).x=g(x)\\] \nis well-defined and free. \nBy Corollary~\\ref{c:hoelder-s1}, this free action extends to a free action \\[\\rho\\co G\/G_0\\to \\Homeo^+(S^1)\\] such that $\\rot\\circ\\rho$ is an injective homomorphism.\n\nFor each $g\\in G$, $x\\in X$ and $n\\in\\bZ$, we have\n\\[\\rho\\circ p(g^n)(x)=g^n(x).\\]\nBy Lemma~\\ref{l:inv} (\\ref{p:inv}) we see that\n\\[\\rot\\circ\\rho\\circ p = \\rot\\restriction_G,\\]\nand hence, that $\\rot\\restriction_G$ is a group homomorphism.\nAs $S^1$ is abelian, we also obtain \n\\[\n[G,G]\\le \\ker(\\rot\\restriction_G)=G_0.\\qedhere\\]\n\\ep\n\n\n\nNote that a finitely generated subgroup of $\\operatorname{SO}(2,\\bR)$ consisting of elements with rational rotation numbers is necessarily finite. So in Lemma~\\ref{l:circle} (\\ref{p:rot-q}), if $G$ is a finitely generated subgroup of $Z(a)$ then $\\rot(G)$ is a finite subgroup of $S^1\\cong\\operatorname{SO}(2,\\bR)$.\n\n\n\n\n\\begin{rem}\nPart (\\ref{p:hom}) of Lemma~\\ref{l:circle} appears in the unpublished work of Farb and Franks \\cite{FF2001}, on which our argument is based. We included here a detailed, self-contained proof for readers' convenience. We also remark the necessity of the infinite--order hypothesis, which was omitted in~\\cite{FF2001}. For example, let us consider $a,b,c\\in\\Diff^\\infty_+(S^1)$ such that \nfor each $x\\in S^1=\\bR\/\\bZ$ \nwe have\n\\[\na(x)=x+1\/2,\\quad\nb(x)=x+1\/4,\\quad\nc(x+1\/2)=c(x)+1\/2\\]\nand such that\n\\[\nc(0) = 0,\\quad\nc(1\/8) = 1\/4,\\quad\nc(1\/4) = 3\/8.\\]\nThen $b,c\\in Z(a)$ and $(bc)^3(0)=0$. Hence we have\n\\[\\rot (b)+\\rot(c) = 1\/4 + 0 \\ne 1\/3 =\\rot(bc).\\]\n\\end{rem}\n\n\n\\begin{lem}\\label{l:interval}\nLet $a\\in \\Diffb(I)$,\nand let $Z(a)$\nbe the centralizer of $a$ in $\\Diffb(I)$.\nThen we have\n\\[\\supp a\\cap \\supp [Z(a),Z(a)]=\\varnothing.\\]\n\\end{lem}\nThe proof is almost identical to that of Lemma~\\ref{l:circle} (\\ref{p:zz}).\n\n\n\n\n\\subsection{The Two--jumps Lemma}\nThe Two--jumps Lemma was developed by Baik and the authors in~\\cite{BKK2016} and is the second essential analytic result needed to establish Theorem~\\ref{thm:main}.\n\n\n\\begin{lem}[Two--jumps Lemma,~\\cite{BKK2016}]\\label{l:fg}\nLet $M\\in\\{I,S^1\\}$ and let $f,g\\co M\\to M$ be continuous maps.\nSuppose $(s_i), (t_i)$ and $(y_i)$ are infinite sequences of points in $M$\nsuch that for each $i\\ge1$, one of the following two conditions hold:\n\\be[(i)]\n\\item\n$f(y_i)\\le s_i = g(s_i) < y_i < t_i = f(t_i) \\le g(y_i)$;\n\\item\n$g(y_i)\\le t_i = f(t_i) < y_i < s_i = g(s_i) \\le f(y_i)$.\n\\ee\nIf $|g(y_i)-f(y_i)|$ converges to $0$ as $i$ goes to infinity,\nthen $f$ or $g$ fails to be $C^1$.\n\\end{lem}\n\nFigure~\\ref{f:fg} illustrates the case (i) of Lemma~\\ref{l:fg}.\nThe reader may note that the homeomorphisms $f$ and $g$ above are \\emph{crossed elements}~\\cite[Definition 2.2.43]{Navas2011}.\nIndeed, the Two--jumps Lemma generalizes a unpublished lemma of Bonatti--Crovisier--Wilkinson regarding crossed $C^1$--diffeomorphisms,\nwhich can be found in \\cite[Proposition 4.2.25]{Navas2011}.\n\n\\begin{figure}[h!]\n \\tikzstyle {bv}=[black,draw,shape=circle,fill=black,inner sep=1pt]\n\\begin{tikzpicture}[>=stealth',auto,node distance=3cm, thick]\n\\draw (-4,0) node (1) [bv] {} node [below] {\\small $f(y_i)$} \n-- (-2,0) node (2) [bv] {} node [below] {\\small $s_i$}\n-- (0,0) node (3) [bv] {} node [below] {\\small $y_i$} \n-- (2,0) node (4) [bv] {} node [below] {\\small $t_i$}\n-- (4,0) node (5) [bv] {} node [below] {\\small $g(y_i)$};\n\\path (3) edge [->,bend right,red] node {} (1);\n\\path (3) edge [->>,bend right,blue] node {} (5);\n\\draw [->>, blue] (2) edge [out = 200,in=-20,looseness=50] (2);\n\\draw [->, red] (4) edge [out = 20,in=160,looseness=50] (4);\n\\end{tikzpicture}%\n\\caption{Two--jumps Lemma.}\n\\label{f:fg}\n\\end{figure}\n\n\n\n\n\\section{The $C^1$--Smoothability of $\\bZ^2\\ast\\bZ$}\nThis section establishes the main technical result of the paper as below.\n\n\n\\begin{thm}\\label{t:tech-main}\nLet $M\\in\\{I,S^1\\}$,\nand let\n $a,b,t\\in\\Diff^1_+(M)$. If \n\\[\\supp a\\cap\\supp b=\\varnothing,\\] \nthen the group $\\form{a,b,t}$ is not isomorphic to $\\bZ^2\\ast\\bZ$.\n\\end{thm}\n\n\n\\begin{rem}\n\\be\n\\item\nWe emphasize that this theorem is about $C^1$, rather than $\\Cbv$, diffeomorphisms. \n\\item The regularity hypothesis of $C^1$ cannot be replaced by $C^0$; see Proposition~\\ref{prop:ab-homeo}.\n\\item The proof of Theorem~\\ref{t:tech-main} is relatively easy and standard if $\\supp a$ or $\\supp b$ is assumed to have finitely many components.\n\\ee\n\\end{rem}\n\nLet us prove Theorem~\\ref{t:tech-main} through a sequence of lemmas in this section.\n\n\\subsection{Finding a lamplighter group from compact support}\nA crucial step in the proof of Theorem~\\ref{t:tech-main} is the following construction,\nwhich generalizes a result of Brin and Squier in the PL setting~\\cite{BS1985}.\nThe same idea to find vanishing words from successive commutators goes back even to the Zassenhaus Lemma~\\cite{Raghunathan1972};\nthe authors thank an anonymous referee for suggesting us to further prove the existence of a lamplighter subgroup.\n\n\\begin{lem}\\label{l:recursive}\nLet $1\\ne g\\in H\\le\\Homeo^+(I)$. If the closure of $\\supp g$ is contained in $\\supp H$, then $H$ contains the lamplighter group $\\bZ\\wr\\bZ$.\n\\end{lem}\n\nMore precisely, we will show that for $g_1=g$, there exists a positive integer $m$ and elements $u_1,\\ldots,u_m\\in H$ such that the recursively defined sequence \\[ g_{i+1}=[g_i,u_i g_i u_i^{-1}],\\quad i=1,2,\\ldots,m,\\] satisfies that\n$\\form{g_m,u_m}\\cong\\bZ\\wr\\bZ$\nand that $g_{m+1}=1$.\nHere and throughout this paper, when $1$ refers to a homeomorphism or a group element then it means the identity, and otherwise it refers to the real number $1$.\n\n\n\n\\bp[Proof of Lemma~\\ref{l:recursive}]\nSince $\\overline{\\supp g_1}$ is a compact subset of the open set $\\supp H$, we can enumerate\n\n\\[ I_1,I_2,\\ldots,I_N\\in\\pi_0(\\supp H)\\]\nsuch that $\\supp g_1\\cap I_i\\ne\\varnothing$ for each $i$ and such that\n\n \\[\\overline{\\supp g_1}\\sse \\bigcup_{i=1}^N I_i.\\]\nNote that each $I_i$ is an open, $H$--invariant interval contained in $(0,1)$.\n\nLet us inductively construct the elements $u_1,\\ldots,u_{k-1}$ satisfying the required properties.\nAs a base case, we put $r(1)=1\\in\\bN$ and \n\\[ K_1 :=\\overline{\\supp g_1}\\cap I_1.\\]\nSince $K_1$ is a nonempty compact subset of $I_1\\sse\\supp H$, we have that\n\\[\n\\sup\\{ u(\\inf K_1)\\mid u\\in H\\}=\\sup I_1>\\sup K_1.\\]\nSo, there exists $u_1\\in H$ such that $K_1\\cap u_1^j K_1=\\varnothing$ for all $j\\in\\bN$. Note that \n\\[\\form{g_1,u_1}\\restriction_{I_1}\\cong\\form{s,t\\mid \\left[s,t^j st^{-j}\\right]=1\\text{ for all }j\\in\\bN}=\\bZ\\wr\\bZ.\\]\nLet us set \\[ r(2) = 1+\\sup\\{ s\\in [1,N]\\mid \\supp [g_1, u_1^j g_1 u_1^{-j}]\\restriction_{I_s}=1\\text{ for all }j\\in\\bN\\}\\ge 1+r(1)=2.\\]\nIf $r(2)>N$, then we have a sequence of surjections\n\\[\n\\bZ\\wr\\bZ\\twoheadrightarrow\n\\form{g_1,u_1}\\twoheadrightarrow\n\\form{g_1,u_1}\\restriction_{I_1}\\twoheadrightarrow\\bZ\\wr\\bZ,\n\\]\nwhich composes to the identity. In particular, $\\form{g_1,u_1}\\cong\\bZ\\wr\\bZ$.\nIn the case where $r(2)\\le N$, we pick $j\\in\\bN$ such that the element\n\\[g_2:= \\left[ g_1, u_1^j g_1 u^{-j}\\right]\\]\nsatisfies $\\supp g_2\\cap I_{r(2)}\\ne\\varnothing$,\nand apply the same argument to $g_2$.\n\nBy a straightforward induction, we eventually find $1\\le m\\le r(m)\\le N$\nand $g_m, u_m\\in G$ such that the following hold:\n\\begin{align*}\n&\\overline{\\supp g_m} \\sse I_{r(m)}\\cup\\cdots\\cup I_N,\\\\\n&\\supp g_m\\cap I_{r(m)}\\ne\\varnothing,\\\\\n&\\supp g_m\\cap u_m^j \\supp g_m\\cap I_{r(m)}=\\varnothing,\\quad\\text{ for all }j\\in\\bN,\\\\\n&\\left[g_m, u_m^j g_m u_m^{-j}\\right] =1\\quad\\text{ for all }j\\in\\bN.\n\\end{align*}\nIt follows that $\\form{g_m,u_m}\\cong \\bZ\\wr\\bZ$.\n\\ep\n\n\n\nLemma~\\ref{l:recursive} implies the following for circle homeomorphisms.\n\n\\begin{lem}\\label{l:global}\nLet $a,b,c,d\\in\\Homeo^+(S^1)$ be nontrivial elements such that\n\\[\n\\supp a \\cap \\supp b = \\varnothing\\quad\\text{ and }\\quad\n\\supp c \\cap \\supp d = \\varnothing.\\]\nIf $\\supp G= S^1$, then $G$ contains $\\bZ\\wr\\bZ$.\n\\end{lem}\n\n\\bp\nFor simplicity, let us abbreviate\n$\\AA = \\pi_0\\supp a$, and similarly define \n$\\BB$, $\\CC$ and $\\DD$.\nSince $S^1$ is compact, there exists a finite open covering $\\mathcal{V}$ of $S^1$ such that \n\\[\n\\mathcal{V}\\sse\\AA\\cup \\BB\\cup\\CC\\cup\\DD.\\]\n\nBy minimizing the cardinality, we can require that $\\mathcal{V}$ forms a \\emph{chain of intervals}.\nMore precisely, this means that\n$\\mathcal{V}=\\{V_1,\\ldots,V_k\\}$ for some $k\\ge1$\nand that \\[\\inf V_i<\\sup V_{i-1}\\le\\inf V_{i+1}\\]\nfor $i=1,2,\\ldots,k$, where here the indices are taken cyclically.\n\nWithout loss of generality, let us assume $V_1\\in\\AA $.\nThen we have $V_{2i-1}\\in\\AA \\cup \\BB$\nand $V_{2i}\\in\\CC\\cup\\DD$ for each $i$. Note that $k$ is an even number\nand that $x=\\inf V_2$ is a global fixed point of $H=\\form{b,c,d}$.\nIn particular, we can regard $H$ as acting on $I$, which is a two-point compactification of $S^1\\setminus \\{x\\}$.\nNote that\n\\[\n\\varnothing\\ne \\overline{\\supp b}\\sse S^1\n\\setminus \\bigcup(\\mathcal{V}\\cap\\AA ) \\sse \\bigcup(\\mathcal{V}\\cap(\\BB\\cup\\CC\\cup\\DD))\n\\sse\\supp H.\\]\nThe desired conclusion follows from Lemma~\\ref{l:recursive}.\n\\ep\n\n\n\\subsection{Supports of commutators}\nWe will need rather technical estimates of supports as given in this subsection.\nIn order to prevent obfuscation of the ideas, we have included some intuition behind the proofs when appropriate. \n\\begin{lem}\\label{l:comm-supp}\nIf $f$ and $g$ are homeomorphisms of a topological space $X$, then\n\\[\\overline{\\supp[f,g]}\\sse \\supp f \\cup \\supp g \\cup \\overline{\\supp f\\cap\\supp g}.\\]\n\\end{lem}\n\\bp\nSuppose \n\\[x\\not\\in \\supp f\\cup \\supp g\\cup \\overline{\\supp f\\cap \\supp g}.\\]\nThen $f(x)=x=g(x)$. Moreover, for some open neighborhood $U$ of $x$ we have \n\\[U\\cap \\supp f\\cap \\supp g =\\varnothing.\\]\nWe can find an open neighborhood $V\\sse U$ of $x$ such that\n\\[f^{\\pm1}(V)\\cup g^{\\pm1}(V)\\sse U.\\]\nLet $y\\in V$. We see $[f,g](y)=y$, by considering the following three cases separately:\n\\[\ny\\in V\\cap \\supp f,\\quad y\\in V\\cap \\supp g,\\quad y\\in V\\cap\\Fix f\\cap \\Fix g.\\]\nSo we obtain that\n\\[[f,g]\\restriction_V=1.\\]\nThis implies \\[x\\not\\in\\overline{\\supp [f,g]}.\\qedhere\\]\n\\ep\n\n\\begin{lem}\\label{lem:c0} \nLet $X$ be a topological space.\nIf $b,c,d\\in\\Homeo(X)$ satisfy \\[\\supp c\\cap\\supp d= \\varnothing,\\]\nthen for $\\phi=[c,bdb^{-1}]$ we have that\n\\[\n\\supp\\phi \\sse \\supp b \\cup cb(\\supp b \\cap\\supp d)\\cup db^{-1}(\\supp b \\cap \\supp c).\\]\n\\end{lem}\n\nLet us briefly explain the key idea behind the statement of this lemma. The support of the homeomorphism $bdb^{-1}$ is exactly $b(\\supp d)$. The homeomorphism $\\phi$ may be viewed as a composition of $bd^{-1}b^{-1}$ and the conjugate of $bdb^{-1}$ by $c$, and the latter of these has support $cb(\\supp d)$. \nBy exhaustively checking the possible images of points under $bdb^{-1}$ and $cbdb^{-1}c^{-1}$, we see that every $x\\in\\supp \\phi$ belongs to one of the three sets as stated in the lemma.\n\n\\bp[Proof of Lemma~\\ref{lem:c0}]\nFor brevity, let us write \n\\[\n\\tilde b = \\supp b,\\quad\n\\tilde c = \\supp c,\\quad\n\\tilde d = \\supp d.\\]\nLet us consider three equivalent expressions for $\\phi$:\n\\[\n [c,bdb^{-1}]=cb d (cb)^{-1}\\cdot bd^{-1}b^{-1}= c\\cdot b\\cdot db^{-1}c^{-1}(db^{-1})^{-1}\\cdot b^{-1}.\\]\nAfter some set theoretic computation, one sees the following.\n\\begin{align*}\\supp \\phi&\\sse\\left(\\tilde c\\cup b\\tilde d\\right)\\cap \\left(cb \\tilde d\\cup b\\tilde d\\right)\\cap\\left(\\tilde c\\cup \\tilde b\\cup db^{-1}\\tilde c\\right)\\\\&\\sse\\left((\\tilde c\\cap cb \\tilde d)\\cup b\\tilde d\\right)\\cap\\left(\\tilde c\\cup \\tilde b\\cup db^{-1}\\tilde c\\right)\\\\&\\sse(\\tilde c\\cap cb \\tilde d) \\cup \\left( b\\tilde d\\cap (\\tilde c\\cup \\tilde b\\cup db^{-1}\\tilde c)\\right)\\\\&\\sse\\left(\\tilde c\\cap cb \\tilde d\\right) \\cup \\left( (\\tilde b\\cup \\tilde d)\\cap (\\tilde b\\cup\\tilde c\\cup db^{-1}\\tilde c) \\right)\\\\ &\\sse\\left(\\tilde c\\cap cb \\tilde d\\right)\\cup\\tilde b\\cup \\left(\\tilde d\\cap (\\tilde c\\cup db^{-1}\\tilde c)\\right)\\\\ &\\sse\\left(\\tilde c\\cap cb \\tilde d\\right)\\cup\\tilde b \\cup\\left(\\tilde d\\cap db^{-1}\\tilde c\\right). \\end{align*}\nNote that we used $b\\tilde d\\sse \\tilde b\\cup \\tilde d$, and also $\\tilde c\\cap\\tilde d=\\varnothing$.\nIt now suffices for us to prove the following claim:\n\\begin{claim*}\\label{claim:cbbd}\nWe have the following:\n\\begin{align*}\n\\tilde c\\cap cb\\tilde d&\\sse cb\\left(\\tilde b\\cap\\tilde d\\right),\\\\\n\\tilde d\\cap db^{-1}\\tilde c&\\sse db^{-1}\\left(\\tilde b\\cap\\tilde c\\right).\n\\end{align*}\n\\end{claim*}\nTo see the first part of the claim, let us consider \n$x\\in X$ satisfying\n\\[cb(x)\\in \\tilde c\\cap cb \\tilde d.\\]\nThen we have $x\\in \\tilde d$ and $cb(x)\\in \\tilde c$.\nSince\n $\\tilde c\\cap \\tilde d=\\varnothing$\nand $b(x)\\in c^{-1}\\tilde c=\\tilde c$, we see $x\\ne b(x)$.\nIn particular, we have $x\\in\\tilde b$\n and $cb(x)\\in cb(\\tilde b\\cap \\tilde d)$. This proves the first part of the claim. The second part follows by symmetry.\\ep\n \n\\begin{lem}\\label{lem:c1} \nIf $b,c,d\\in\\Diff^1_+(I)$ are given such that \\[\\supp c\\cap\\supp d= \\varnothing,\\]\nthen for $\\phi=[c,bdb^{-1}]$ we have that\n\\[\\overline{\\supp\\phi\\setminus \\supp b}\\sse\\supp c\\cup\\supp d.\\]\n\\end{lem}\n\\bp\nAs in the proof of Lemma~\\ref{l:global}, we let $\\BB=\\pi_0\\supp b$ and $\\CC=\\pi_0\\supp c$.\nLet\n\\[J_B=B\\cup cb(B\\cap\\supp d)\\cup db^{-1}(B\\cap \\supp c)\\]\nfor each $B\\in\\BB$.\nBy Lemma~\\ref{lem:c0}, we have that\n\\[\\supp\\phi\\sse \\bigcup\\{J_B\\mid B\\in\\BB\\}\n= \\bigcup\\{J_B\\setminus B\\mid B\\in \\BB\\}\\cup \\supp b.\\]\nMoreover, for each $B\\in\\BB$ we note that\n\\[\n\\overline{\nJ_B\\setminus B}\n\\sse \\overline{c(B)\\setminus B}\\cup \\overline{d(B)\\setminus B}\n\\sse\\supp c \\cup\\supp d.\\]\n\n\\begin{claim*}\\label{claim:jbb}\nThe following set is a finite collection of intervals:\n\\[ \\BB_0=\\{B\\in\\BB\\mid J_B\\ne B\\}.\\]\n\\end{claim*}\nWe will employ the $C^1$--hypothesis for this claim. \nLet us write\n\\[\\BB_1=\\{B\\in\\BB\\mid cb(B\\cap \\supp d)\\setminus B\\ne\\varnothing\\},\\ \n\\BB_2=\\{B\\in\\BB\\mid db^{-1}(B\\cap \\supp c)\\setminus B\\ne\\varnothing\\}.\\]\nAssume for a contradiction that $\\BB_0=\\BB_1\\cup\\BB_2$ is infinite. \nWe may suppose $\\BB_1$ is infinite, as the proof is similar when $\\BB_2$ is infinite.\nThere are infinitely many distinct $B_1,B_2,\\ldots\\in \\BB_1$ \nand $x_i\\in B_i\\cap\\supp d$ such that $cb (x_i)\\not\\in B_i$. \nThen we have $C_1,C_2,\\ldots\\in \\CC$ such that\n$b (x_i), cb(x_i)\\in C_i$.\nSince $x_i\\in\\supp d$, we have $x_i\\not\\in C_i$; see Figure~\\ref{f:JB}.\n\nLet us consider the interval $J_i = [x_i,cb (x_i)]$ which contains $b(x_i)$ in the interior, up to switching the endpoints of this interval.\nThen we have\n\\[(b(x_i),cb (x_i)]\\cap \\partial B_i\\ne\\varnothing,\\quad [x_i,b (x_i))\\cap\\partial C_i\\ne\\varnothing.\\]\nWe now apply the Two--jumps Lemma (Lemma~\\ref{l:fg}) to the following parameters\n\\[f=b^{-1},\\ g=c,\\ s_i=\\partial C_i\\cap B_i,\\ t_i=\\partial B_i\\cap C_i,\\ y_i=b(x_i).\\]\nWe deduce that $b$ or $c$ is not $C^1$.\nThis is a contradiction and the claim is proved.\n\nFrom the claim above, we deduce the conclusion as follows.\n\\[\n\\overline{\\supp\\phi\\setminus\\supp b}\n\\sse\\overline{\\bigcup\\{J_B\\setminus B\\mid B\\in\\BB_0\\}}=\\bigcup\\{\\overline{J_B\\setminus B}\\mid B\\in\\BB_0\\}\n\\sse\\supp c\\cup\\supp d.\\qedhere\\]\n\\ep\n\n\\begin{figure}[h!]\n \\tikzstyle {a}=[black,postaction=decorate,decoration={%\n markings,%\n mark=at position 1 with {\\arrow[black]{stealth};} }]\n \\tikzstyle {bv}=[black,draw,shape=circle,fill=black,inner sep=1.5pt]\n{\n\\begin{tikzpicture}[thick,scale=.7]\n\\draw [red,ultra thick] (-4,0.5) -- (2,0.5);\n\\draw [blue,ultra thick] (0,0) -- (6,0);\n\\draw [teal,ultra thick] (4,.5) -- (10,.5);\n\\draw (-3,0) node [] {\\small $\\supp d$}; \n\\draw (9,0) node [] {\\small $C_i$}; \n\\draw (3,-.5) node [] {\\small $B_i$}; \n\\draw [dashed] (1,1.4) -- (1,-.3) node [below] {\\small $x_i$};\n\\draw [dashed] (5,1.4) -- (5,-.3) node [below] {\\small $b(x_i)$};\n\\draw [dashed] (7.5,1.4) -- (7.5,-.3) node [below] {\\small $cb(x_i)$};\n\\draw (1,1) -- (7.5,1);\n\\draw (3,1.4) node [] {\\small $J_i$}; \n\\end{tikzpicture}%\n}\n\\caption{Lemma~\\ref{lem:c1}.}\n\\label{f:JB}\n\\end{figure}\n\n\\subsection{Finding compact supports}\nWe will deduce Theorem~\\ref{t:tech-main} from the following, seemingly weaker result.\n\\begin{lem}\\label{l:main}\nLet $M\\in\\{I,S^1\\}$ and let\n$a,b,c,d\\in\\Diff^1_+(M)$.\nIf \n\\[\n\\supp a\\cap\\supp b=\\varnothing,\\quad\n\\supp c\\cap\\supp d=\\varnothing,\\]\nthen the group $\\form{a,b,c,d}$ is not isomorphic to $\\bZ^2\\ast\\bZ^2$.\n\\end{lem}\n\nLet us note two properties of RAAGs. First, a RAAG does not contain a subgroup isomorphic to $\\bZ\\wr\\bZ$. The reason is that, every two--generator subgroup of a RAAG is either free or free abelian~\\cite{Baudisch1981}; see also~\\cite[Corollary 1.3]{KK2015GT}.\nSecond, a RAAG is \\emph{Hopfian}; that is, every endomorphsm of a RAAG is an isomorphism. This follows from a general fact that every finitely generated residually finite group is Hopfian~\\cite{LS2001}.\n\n\n\\bp[Proof of Theorem~\\ref{t:tech-main} from Lemma~\\ref{l:main}]\nAssume $\\form{a,b,t}\\cong\\bZ^2\\ast\\bZ$. \nSince the RAAG $\\bZ^2\\ast\\bZ$ is Hopfian, the natural surjection between groups\n\\[ \\form{A,B,T\\mid [A,B]=1}\\to \\form{a,b,t}\\]\nis actually an isomorphism. It follows that\n\\[\\form{a,b,tat^{-1},tbt^{-1}}\\cong\\form{A,B,TAT^{-1},TBT^{-1}}\\cong\\bZ^2\\ast\\bZ^2.\\]\nThis contradicts Lemma~\\ref{l:main}, since the four diffeomorphism $a,b,tat^{-1},tbt^{-1}$ satisfy the conditions of the lemma.\n\\ep\n\n\\bp[Proof of Lemma~\\ref{l:main}]\nWe put $G=\\form{a,b,c,d}$ and consider an abstract group\n\\[ G_0 = \\form{a_0,b_0,c_0,d_0\\mid [a_0,b_0]=1=[c_0,d_0]}\\cong \\bZ^2\\ast\\bZ^2.\\]\nThere is a natural surjection $p\\co G_0\\to G$ defined by \\[(a_0,b_0,c_0,d_0)\\mapsto(a,b,c,d).\\] \n\nAssume for a contradiction that $G\\cong G_0$. By the Hopficity of $\\bZ^2\\ast\\bZ^2$, we see that $p$ is an isomorphism. Since $G$ does not contain $\\bZ\\wr\\bZ$, Lemma~\\ref{l:global} implies that $G$ has a global fixed point.\nIn other words, we may assume $M=I$.\n\nLet us define \n$\\phi=[c,bdb^{-1}]$ and $\\psi=[\\phi,a]$.\nLemma~\\ref{l:comm-supp} implies that\n\\[\\overline{\\supp\\psi}\\sse\\supp\\phi\\cup\\supp a\\cup\\overline{\\supp\\phi\\cap\\supp a}.\\]\nWe see from Lemma~\\ref{lem:c1} that\n\\[\\overline{\\supp\\phi\\cap\\supp a}\n\\sse\\overline{\\supp\\phi\\setminus\\supp b}\n\\sse\\supp c\\cup \\supp d.\\]\nSo, it follows that\n\\[\n\\overline{\\supp \\psi}\\sse\\supp G.\\]\n\nAs we are assuming $p$ is injective, we have \n\\[\\psi = [\\phi,a]=\\left[ [c, bdb^{-1}],a\\right]\\ne1.\\]\nLemma~\\ref{l:recursive} implies that $G$ contains $\\bZ\\wr\\bZ$, which is a contradiction. This completes the proof.\\ep\n\nLet us conclude this section by describing one generalization of Theorem~\\ref{t:tech-main}.\n\n\\begin{prop}\\label{t:tech-main2}\nLet $M\\in\\{I,S^1\\}$.\nIf $a,b,c,d\\in\\Diff^1_+(M)$ satisfy \n\\[\\supp a\\cap\\supp b=\\varnothing,\\quad\n\\supp c\\cap\\supp d=\\varnothing.\\]\nand \n\\[\n\\left[[c,bdb^{-1}],a\\right]\\ne1,\n\\]\nthen $\\form{a,b,c,d}$ contains the lamplighter group $\\bZ\\wr\\bZ$.\n\\end{prop}\n\nIn particular, the group $\\form{a,b,c,d}$ does not embed into a RAAG.\n\n\n\n\n\n\n\n\\section{Proof of Theorem~\\ref{thm:main}}\\label{s:main-thm}\n\nIn this section, we apply the facts we have gathered to complete the proof of the main result.\n\n\\subsection{Reducing to the connected case}\nWe will reduce the proof of Theorem~\\ref{thm:main} to the case $M\\in\\{I,S^1\\}$, using the following group theoretic observations.\n\n\\begin{lem}\\label{lem:L1}\nSuppose $A,B,C,D$ are groups, and suppose that $A\\times B$ is a normal subgroup of $C*D$. Then at least one of these four groups is trivial.\n\\end{lem}\n\\begin{proof}\nIf $A\\times B$ is a subgroup of $C*D$ then the Kurosh Subgroup Theorem implies that there is a free product decomposition \\[A\\times B\\cong F\\ast\\bigast_{i} H_i,\\] where $F$ is a free group (possibly of infinite rank) and where each $H_i$ is conjugate into $C$ or into $D$. By analyzing centralizers of elements, it is easy to show that a nontrivial free product is never isomorphic to a nontrivial direct product (cf.~\\cite[p.177]{LS2001}). It follows that $A\\times B$ is conjugate into $C$ or $D$, which contradicts the normality of $A\\times B$.\n\\end{proof}\n\nAn alternative proof of Lemma~\\ref{lem:L1} can be given using Bass--Serre theory (see~\\cite{Serre1977}).\n\n\\begin{lem}\\label{lem:L2}\nSuppose $A,B,C,D$ are nontrivial groups, and that $A*B\\le C\\times D$. Then there is an injective homomorphism from $A*B$ into either $C$ or $D$.\n\\end{lem}\n\\begin{proof}\nSuppose the contrary, so that $K_C$ and $K_D$ are the (nontrivial) kernels of the inclusion of $A*B$ into $C\\times D$ composed with the projections onto $C$ and $D$. Then $K_C\\cap K_D=1$ and $K_C$ and $K_D$ normalize each other, so that $K_CK_D\\cong K_C\\times K_D\\le A*B$. This contradicts Lemma~\\ref{lem:L1}.\n\\end{proof}\n\n\n\n\n\n\n\n\\begin{lem}\\label{lem:connected}\nSuppose \\[M=\\coprod_{i=1}^n M_i\\] is a compact one--manifold, and suppose that $A\\ast B$ embeds into $\\Diffb(M)$. Then \n for some finite index subgroups $A_0\\le A$ and $B_0\\le B$,\n and for some $i$, we have an embedding of $A_0\\ast B_0$ into $\\Diffb(M_i)$.\n\\end{lem}\n\\begin{proof}\nThis follows immediately from Lemma~\\ref{lem:L2}, using the fact that $\\Diffb(M)$ is commensurable with \\[\\prod_{i=1}^n\\Diffb(M_i).\\]\n\nNote that passage to finite index subgroups is necessary, since $M$ may consist of a union of diffeomorphic manifolds which are permuted by the action of $A\\ast B$.\n\\end{proof}\n\n\\subsection{Taming supports}\n\nLet us denote the center of a group $G$ by $Z_G$.\nIf $M$ is a one--manifold and if $s\\in\\Diffb(M)$, then\nwe denote by $Z(s)$ the centralizer of $s$ in $\\Diffb(M)$.\nThe following lemma is crucial for applying Lemma~\\ref{l:main}:\n\n\n\\begin{lem}\\label{lem:Z2}\nAssume one of the following:\n\\be[(i)]\n\\item\n$M=I$ and $G$ is a nonabelian group such that $Z_G\\ne1$.\n\\item\n$M=S^1$ and $G$ is a non-metabelian group such that $\\bZ\\le Z_G$.\n\\item\n$M=S^1$ and $G$ is a finitely generated group such that $\\bZ\\le Z_G$\nand such that $G$ is not abelian-by-finite cyclic.\n\\ee\nIn each of the cases, if $G\\le\\Diffb(M)$, then there is a subgroup $\\bZ^2\\le G$ generated by diffeomorphisms $a$ and $b$ such that $\\supp a\\cap\\supp b=\\varnothing$.\n\\end{lem}\n\n\n\\bp\n\n\\emph{Case (i).} Let us pick $s\\in Z_G\\setminus1$ and $b\\in[G,G]\\setminus1$. \nSince $G\\le Z(s)$, Lemma~\\ref{l:interval} implies that \n\\[\\supp s\\cap \\supp b=\\varnothing.\\]\nSince $\\Homeo^+(I)$ is torsion-free, we have $\\form{b,s}\\cong\\bZ^2$ as desired.\n\n\\emph{Case (ii).}\nWe are given with some $s\\in Z_G$ such that $\\form{s}\\cong\\bZ$. As $G$ is nonabelian, Lemma~\\ref{l:circle} (\\ref{p:rot-irr}) implies that $\\rot(s)\\in\\bQ$; in particular, $s^n$ is grounded for some $n\\ge1$.\nFrom part (\\ref{p:hom}) of the same lemma\nand from that $G\\le Z(s)$, we see \nevery element of $[G,G]$ is grounded.\nSince $[G,G]\\le Z(s^n)$, we note from Lemma~\\ref{l:circle} (\\ref{p:zz}) that \n\\[\\supp(s^n)\\cap \\supp G''=\\varnothing.\\]\nFrom the metabelian hypothesis, we can find $b\\in G''\\setminus1$.\nAs $s^n$ and $b$ are grounded, they have infinite orders. It follows that $\\form{s^n,b}\\cong\\bZ^2$.\n\n\n\\emph{Case (iii).}\nLet us proceed similarly to the case (ii). Namely, pick $s\\in Z_G$ such that $\\form{s}\\cong\\bZ$. \nBy Lemma~\\ref{l:circle}, we have a homomorphism\n\\[\n\\rot\\restriction_G\\co G\\to \\bQ.\\]\nWe fix $n\\ge1$ such that $s^n$ is grounded.\nAs $G$ is finitely generated, we see $\\rot(G)$ is finite cyclic.\nThe hypothesis implies that $G_0=\\ker(\\rot\\restriction_G)$ is not abelian.\nSince every element of $G_0$ is grounded, we can apply\n Lemma~\\ref{l:circle} (\\ref{p:zz}) and deduce \n\\[\\supp s^n\\cap \\supp [G_0,G_0]=\\varnothing.\\]\nEach $b\\in [G_0,G_0]\\setminus1$ then yields the desired subgroup\n$\\form{b,s^n}\\cong\\bZ^2$.\n\\ep\n\n\\subsection{The main result}\n\n\n\\begin{proof}[Proof of Theorem~\\ref{thm:main}]\nSuppose $(G\\times\\Z)*\\Z\\le \\Diffb(M)$ for some compact one--manifold $M$. Replacing $G$ by a finite index subgroup if necessary, we may assume that $M$ is connected, by Lemma~\\ref{lem:connected}. \nBy applying the cases (i) and (ii) of Lemma~\\ref{lem:Z2} to the group $G\\times \\bZ$,\nwe can find a subgroup $\\langle a,b\\rangle\\cong\\bZ^2\\le G\\times\\Z$ such that $\\supp a\\cap\\supp b=\\varnothing$.\nIf we write the $\\bZ$--free factor of $(G\\times \\bZ)\\ast\\bZ$ as $\\form{t}$, then\n\\[\\form{a,b,t}\\cong\\form{a,b}\\ast\\form{t}\\cong\\bZ^2\\ast\\bZ.\\]\nThis contradicts Theorem~\\ref{t:tech-main}.\n\\end{proof}\n\nOne can now deduce Corollary~\\ref{cor:fg1} as well as Corollary~\\ref{cor:cnt} below from Lemma~\\ref{lem:Z2}, in the exact same fashion as Theorem~\\ref{thm:main}. \n\\begin{cor}\\label{cor:cnt}\nLet $G$ be a group.\n\\be\n\\item \nIf $G$ is nonabelian and if the center of $G$ is nontrivial,\nthen\n$G*\\Z$ admits no faithful $\\Cb$ action on $I$. \n\\item Suppose $G$ is finitely generated.\nIf $G$ is not abelian-by-finite cyclic\nand if the center of $G$ contains a copy of $\\bZ$,\nthen \n$G*\\Z$ admits no faithful $\\Cb$ action on $S^1$. \n\\end{enumerate}\n\\end{cor}\n\n\n\nHere, a group $G$ is $\\mathcal{X}$--by--$\\mathcal{Y}$ for group theoretic properties $\\mathcal{X}$ and $\\mathcal{Y}$ if there is an exact sequence \\[1\\to K\\to G\\to Q\\to 1\\] such that $K$ has property $\\mathcal{X}$ and $Q$ has property $\\mathcal{Y}$. We allow both $K$ and $Q$ to be trivial.\n\nNote that $G\\times \\bZ$ often occurs as a subgroup of $\\Diffb(M)$, where $G$ is not virtually metabelian. We have the following immediate consequence:\n\n\\begin{cor}\\label{cor:not closed}\nLet $\\mathcal{G}$ denote the class of finitely generated subgroups of $\\Diffb(M)$. The class $\\mathcal{G}$ is not closed under taking finite free products.\n\\end{cor}\n\n\n\n\n\\section{Smooth right-angled Artin group actions on compact one--manifolds}\n\nIn this and the remaining sections, we deduce several corollaries from Theorem \\ref{thm:main}. We first complete the classification of right-angled Artin groups which admit faithful $C^{\\infty}$ actions on a compact one--manifold (Corollary \\ref{cor:classification}).\n\n\\begin{lem}\\label{lem:dichotomy}\nLet $A(\\gam)$ be a right-angled Artin group. Then one of the following mutually exclusive conclusions holds:\n\\begin{enumerate}\n\\item\nWe have $(F_2\\times\\Z)*\\Z\\le A(\\gam)$;\n\\item\nThe graph $\\gam$ lies in $\\mathcal{K}_3$.\n\\end{enumerate}\n\\end{lem}\n\\begin{proof}\nLet us consider a stratification of graph classes:\n\\[ \\KK_2\\sse\\KK_3\\sse\\KK.\\]\n\nSuppose $\\gam\\in\\KK_2$. Then $A(\\gam)$ is the free product of free abelian groups, and hence contains no copy of $(F_2\\times\\Z)*\\Z$.\n\nLet $\\gam\\in\\KK_3\\setminus\\KK_2$. Then $\\gam$ is the join of at least two graphs $\\gam_1,\\gam_2$ in $\\KK_2$. \nWe write\n\\[\nA(\\gam)=A(\\gam_1)\\times A(\\gam_2).\\]\nIf $A(\\gam)$ contains a copy of $(F_2\\times\\bZ)\\ast\\bZ$,\nthen so does $A(\\gam_1)$ or $A(\\gam_2)$\nby Lemma~\\ref{lem:L2}; this would contradict the previous paragraph.\n\nAssume $\\gam\\in\\KK\\setminus\\KK_3$.\nFirst consider the case that $\\gam\\in\\KK_{2i}\\setminus\\KK_{2i-1}$ for some $i\\ge2$. \nWe can write\n\\[\n\\gam=\\coprod_{j=1}^k \\gam_j\\]\nfor some $k\\ge2$ and for some nonempty connected graphs $\\gam_j\\in\\KK_{2i-1}$. \nThese graphs $\\Gamma_j$ cannot all be complete graphs, for otherwise $\\gam\\in\\KK_2$. \nSo at least one graph $\\Gamma_j$ contains $P_3$, the path on three vertices, as a full subgraph. \nThis implies that $A(\\gam)$ contains a copy of $(F_2\\times\\bZ)\\ast\\bZ$.\n\nWe then consider the case that $\\gam\\in\\KK_{2i+1}\\setminus\\KK_{2i}$ for some $i\\ge2$. \nNote $\\gam$ is the join of some graphs $\\gam_1,\\ldots,\\gam_k$ in $\\KK_{2i}$. \nBy the previous graph, each $A(\\gam_j)$ contains $(F_2\\times\\bZ)\\ast\\bZ$.\n\nFinally assume $\\gam\\notin\\KK$, so that $\\gam$ is not a cograph. \nThen we have that $P_4$ is a full subgraph of $\\gam$, so that $A(P_4)\\le A(\\gam)$. The group $A(P_4)$ contains every right-angled Artin group $A(F)$, where $F$ is a finite forest (see~\\cite{KK2013}). Since the defining graph of $(F_2\\times\\Z)*\\Z$ is a copy of a path $P_3$ on three vertices together with an isolated vertex, its defining graph is a finite forest. We see that $(F_2\\times\\Z)*\\Z\\le A(\\gam)$.\n\\end{proof}\n\n\nWe complete the proof of Corollary \\ref{cor:classification} with the following proposition:\n\n\\begin{prop}\\label{prop:cinfty}\nLet $\\gam\\in\\mathcal{K}_3$ and let $M$ be a compact one--manifold. Then there is an embedding of $A(\\gam)$ into $\\Diff_+^{\\infty}(M)$.\n\\end{prop}\n\n\nIn the case when $M=S^1$, we will prove more precise facts in Section \\ref{sec:semiconj}.\n\n\\begin{proof}[Proof of Proposition~\\ref{prop:cinfty}]\nLet $\\Diff^{\\infty}_0(I)$ denote the group of $C^{\\infty}$ diffeomorphisms of the interval which are infinitely tangent to the identity at $\\{0,1\\}$. It suffices to find a copy $A(\\gam)\\le\\Diff^{\\infty}_0(I)$, since $I$ is a submanifold of every compact one--manifold $M$ and every element of $\\Diff^{\\infty}_0(I)$ can, by definition, be extended to all of $M$ by the identity map. \n\nIn the case when $\\gam\\in\\mathcal{K}_2$, we can write \n$A(\\gam)=\\Z^{n_1}\\ast\\cdots\\ast\\Z^{n_k}$.\nAs we have an embedding $A(\\gam)\\hookrightarrow\\Z*\\Z^N$ for $N=\\max_i n_i$, it suffices to prove the proposition for $\\Z*\\Z^N$ in this case. \nTo do this, we first find a copy of $\\Z^N\\le \\Diff^{\\infty}_0(I)$ such that the support of each nontrivial element of $\\Z^N$ is all of $(0,1)$. The existence of such a copy of $\\Z^N$ follows from choosing a $C^{\\infty}$ vector field on $I$ which vanishes only at $\\partial I$ and integrating it to get a flow, which gives an $\\R$--worth of commuting elements of $\\Diff^{\\infty}_0(I)$.\nNow, choosing a generic (in the sense of Baire) element $\\psi$ of $\\Diff^{\\infty}_0(I)$, we have that $\\psi$ and this copy of $\\Z^N$ generate a copy of $\\Z*\\Z^N\\le\\Diff^{\\infty}_0(I)$ (cf.~\\cite{KKMj2016}).\n\n\n\nFinally, if $\\gam\\in\\mathcal{K}_3\\setminus\\mathcal{K}_2$ then $A(\\gam)$ is a finite direct product of $k$ right-angled Artin groups with defining graphs in $\\mathcal{K}_2$. Write again $N$ for the maximal rank of an abelian subgroup of $A(\\gam)$. We choose a finite collection of disjoint intervals $\\{J_1,\\ldots,J_k\\}$ with nonempty interior inside of $I$, and realize a copy of $\\Z*\\Z^N$ on each $J_i$, extending by the identity outside of $J_i$. It is clear that $A(\\gam)$ is thus realized as a subgroup of $\\Diff^{\\infty}_0(I)$.\n\\end{proof}\n\n\n\n\n\n\n\n\n\n\\section{Lower regularity}\n\nIn this section, we prove Proposition \\ref{prop:free prod}. \nRecall a \\emph{left order} on a group $G$ is a total order $\\prec$ on $G$ such that for all triples $a,b,g\\in G$ we have $a\\prec b$ if and only if $ga\\prec gb$. \nA group is \\emph{left orderable} if it admits a left order.\n\n\nEvery subgroup of $\\Homeo^+(\\bR)$ is left orderable. Conversely, if $G$ is countable and left orderable, then there is a faithful action $G\\to\\Homeo^+(\\bR)$;\nit can be further required from the action that for some fixed point $x_0\\in\\bR$,\nwhenever $g\\prec h$ we have $g(x_0)1$, we choose points $d_{i-1}}@\/_2pc\/[rrrr] \n& \\ar@{=>}[l] \\lim_{{\\tL},B} q'_{*} p'^{*} & \\ar@{=>}[l] \\lim_{{\\tL},B} p^{*}q_{*} \\ar@{=>}[r] & p^{*}q_{*} \\lim_{{\\tL},C} \\ar@{=>}[r] & q'_{*} p'^{*}\\lim_{{\\tL},C}.}\n\\]\n\\\\\nAll the natural transformations except the curved one on the bottom are invertible. This follows from the base change natural equivalences and the fact that the pushforward functors are right adjoints and so commute with limits. Therefore, the natural transformation on the bottom is also invertible.\nSince $q'_{*}$ is conservative, the natural transformation $\\lim_{{\\tL},C\\otimes_{A}B} p'^{*} \\Longrightarrow p'^{*}\\lim_{{\\tL},C}$ of functors ${\\text{\\bfseries\\sf{Mod}}}(C)^{{\\tL}} \\to {\\text{\\bfseries\\sf{Mod}}}(C \\otimes_{A} B)$ is also invertible and so $p'^{*}$ commutes with finite limits. Therefore, $p'$ is flat.\n\\ \\hfill $\\Box$\n\\begin{lem}\\label{lem:left}Let ${\\tC}$ be a closed symmetric monoidal category with all finite limits and colimits. Let $p$ be a morphism in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ and let $p'$ be a base change of $p$. If $p$ is a monomorphism then so is $p'$. If $p$ is of finite presentation, so is $p'$.\n\\end{lem}\n{\\bf Proof.} Left to the reader.\n\\ \\hfill $\\Box$\n\nConsider the following very slight modification of Proposition 2.4 and its proof from \\cite{TV} (we only require finite limits and colimits and consider only a special case of their proposition).\n\\begin{lem}\\label{lem:BaseChangeConserv}Let ${\\tC}$ be a closed symmetric monoidal category with all finite limits and colimits. Suppose that a family $\\{p_{i}:X_{i}\\to X\\}$ in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ is such that the family $\\{p^{*}_{i}:{\\text{\\bfseries\\sf{Mod}}}(X) \\to {\\text{\\bfseries\\sf{Mod}}}(X_i)\\}$ has a finite conservative subfamily. Then any pull-back family $\\{p_{i}:X_{i}\\times_{X} Y\\to Y\\}$ coming from a base change $Y\\to X$ has the same property.\n\\end{lem}\n{\\bf Proof.} In order to show the base change property, consider $q:Y \\to X$. Choose a finite set $J \\subset I$ such that $\\prod_{i \\in J} p^{*}_{i}$ is conservative. Consider the functor $\\prod_{i \\in J} p^{'*}_{i}$ where $q'_{i}$, $p'_i$ and $p_i$ play the role of $q'$, $p'$ and $p$ in diagram (\\ref{basechange}). In order to show it is conservative, its enough to show that $\\prod_{i \\in J} q'_{i*}p^{'*}_{i}$ is conservative but using equation (\\ref{eqn:basechangeequation}) this is isomorphic to $(\\prod_{i \\in J} p^{*}_{i}) q_{*}$ which is conservative since $q_{*}$ is conservative.\n\\ \\hfill $\\Box$\n\\begin{prop}\\label{TV1}\nLet ${\\tC}$ be a closed symmetric monoidal category with all finite limits and colimits. Consider the families $\\{p_{i}:X_{i}\\to X\\}_{i \\in I}$ in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ such that the family $\\{p^{*}_{i}:{\\text{\\bfseries\\sf{Mod}}}(X) \\to {\\text{\\bfseries\\sf{Mod}}}(X_i)\\}_{i \\in I}$ has a finite conservative subfamily and that each $p_{i}^{*}$ is left exact. These families define a pretopology on ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$. \n\\end{prop}\n{\\bf Proof.} In order to show the base change property, consider $q:Y \\to X$ in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ and let $q'_{i}$, $p'_i$ and $p_i$ play the role of $q'$, $p'$ and $p$ in diagram (\\ref{basechange}). Lemma \\ref{lem:BaseChangeConserv} implies that the family $\\{p^{*}_{i}\\}$ has a finite conservative subfamily. The fact that the $p_{i}^{*}$ are exact follows from Lemma \\ref{InvBaseChange}.\n\n\\ \\hfill $\\Box$\n\\begin{defn}\\label{defn:fpqc}\nFor any closed symmetric monoidal category ${\\tC}$ which has all finite limits and colimits, the topology coming from Proposition \\ref{TV1} is called the fpqc topology on ${\\text{\\bfseries\\sf{Aff}}}({\\tC})= {\\text{\\bfseries\\sf{Comm}}}({\\tC})^{op}$. When equipped with this topology, we denote this category by ${\\text{\\bfseries\\sf{Aff}}}({\\tC})^{fpqc}.$ The category of sheaves of sets is denoted ${\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{Aff}}}({\\tC})^{fpqc}).$\n\\end{defn}\n\\begin{defn}\\label{defn:fTVZ}\n The morphism $\\spec(B) \\to \\spec(A)$ is called a formal Zariski open immersion if the corresponding morphism $A \\to B$ in ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is a flat epimorphism (defined in Remark \\ref{rem:epi} and Definition \\ref{defn:TVf}).\n\\end{defn}\n\\begin{defn}\\label{defn:TVZ}\n The morphism $\\spec(B) \\to \\spec(A)$ is called a Zariski open immersion if the corresponding morphism $A \\to B$ in ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is a flat epimorphism of finite presentation (defined in Remark \\ref{rem:epi} and Definitions \\ref{defn:FinitePres} and \\ref{defn:TVf}).\n\\end{defn}\n\\begin{lem}\\label{lem:ConsAgain}If a family $\\{A\\to B_{i}\\}_{i\\in I}$ is conservative and $A'$ is any $A$-algebra then the family $\\{A'\\to B_{i}\\otimes_{A} A'\\}_{i\\in I}$ is conservative.\n\\end{lem}\n{\\bf Proof.}\nThis has already been shown in Proposition \\ref{TV1}.\n\\ \\hfill $\\Box$\n\n\\begin{prop}\\label{prop:formalZarPretop}\nThere is a pretopology whose covering families $\\{A\\to B_{i}\\}_{i\\in I}$ are those families where each $A\\to B_{i}$ is a formal Zariski open immersion and the family $\\{A\\to B_{i}\\}_{i\\in I}$ has a finite conservative subfamily.\n\\end{prop}\n{\\bf Proof.}\nThis follows immediately from Lemmas \\ref{InvBaseChange}, \\ref{lem:ConsAgain} and \\ref{lem:left}.\n\\ \\hfill $\\Box$\n\\begin{defn}\\label{defn:fTV2} The formal Zariski topology on ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ is the topology associated to the pretopology from Proposition \\ref{prop:formalZarPretop}. When equipped with this topology, we denote the category by ${\\text{\\bfseries\\sf{Aff}}}({\\tC})^{fZar}$. The category of sheaves of sets is denoted ${\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{Aff}}}({\\tC})^{fZar}).$\n\\end{defn}\n\\begin{prop}\\label{prop:ZarPretop}\nThere is a pretopology whose covering families $\\{A\\to B_{i}\\}_{i\\in I}$ are those families where each $A\\to B_{i}$ is a Zariski open immersion and the family $\\{A\\to B_{i}\\}_{i\\in I}$ has a finite conservative subfamily.\n\\end{prop}\n{\\bf Proof.}\nThis follows immediately from Lemmas \\ref{InvBaseChange}, \\ref{lem:ConsAgain} and \\ref{lem:left}.\n\\ \\hfill $\\Box$\n\\begin{defn}\\label{defn:TV2} The Zariski topology on ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ is the topology associated to the pretopology from Proposition \\ref{prop:ZarPretop}. When equipped with this topology, we denote the category by ${\\text{\\bfseries\\sf{Aff}}}({\\tC})^{Zar}$. The category of sheaves of sets is denoted ${\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{Aff}}}({\\tC})^{Zar}).$\n\\end{defn}\n\\begin{defn}\\label{defn:hRep} For any affine object $\\spec(A),$ $A \\in {\\text{\\bfseries\\sf{Comm}}}(\\text{\\bfseries\\sf{C}}),$ the presheaf of sets $\\straighth_{A}$ is given by\n\\[{\\text{\\bfseries\\sf{Aff}}}({\\tC}) \\to {\\text{\\bfseries\\sf{Set}}}\n\\]\n\\[\\spec(B) \\mapsto \\Hom_{{\\text{\\bfseries\\sf{Comm}}}(\\text{\\bfseries\\sf{C}})}(A,B).\n\\]\n\\end{defn}\nCor. 2.11 of \\cite{TV} implies that \n\\begin{prop}\\label{TV3}\nFor any $A \\in {\\text{\\bfseries\\sf{Comm}}}(\\text{\\bfseries\\sf{C}})$, the preseheaf $\\straighth_{A}$ is a sheaf for fpqc, the formal Zariski and the Zariski topologies. \n\\end{prop}\n\n\\begin{defn}\\label{defn: ZarOpenSch}\\cite{TV} Let $X \\in {\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$ and $F\\in {\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{Zar})$ be a subsheaf of $X$. Then $F$ is a called a Zariski open of $X$ if there is a family of Zariski opens $\\{X_i \\to X\\}_{i \\in I}$ such that $F$ is the image of the morphism of sheaves $\\coprod_{i \\in I} X_i \\to X$. A morphism $F \\to G$ in ${\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{Zar})$ is called is a Zariski open immersion if for every $X \\in {\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$ and every $X \\to G$ the induced morphism $F \\times_{G} X \\to X$ is a monomorphism whose image is a Zariski open in $X$.\nThe category ${\\text{\\bfseries\\sf{Sch}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{Zar})$ of schemes is defined to be the full subcategory of ${\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{Zar})$ of sheaves $F$ such that there exists a family of $X_{i} \\in {\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$ for $i \\in I$ and a morphism $p: \\coprod_{i \\in I} X_i \\to F$ such that $p$ is an epimorphism of sheaves and for each $i$ the morphism $X_i \\to F$ is a Zariski open.\n\\end{defn}\n\n\n\\begin{defn}\\label{defn:GeneralSch} Suppose that ${\\text{\\bfseries\\sf{A}}}$ is a full subcategory of ${\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$ and suppose that $\\tau$ is a subcategory of ${\\text{\\bfseries\\sf{A}}}$ with all objects and such that all morphisms in $\\tau$ are monomorphisms and so that the base change of a morphism in $\\tau$ by an arbitrary morphism of ${\\text{\\bfseries\\sf{A}}}$ is in $\\tau$. Say that $T$ is a pre-topology whose covers consist of families of morphisms where each morphism in the cover belongs to $\\tau$. Let $X \\in {\\text{\\bfseries\\sf{A}}}$ and $F\\in {\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{A}}}^{T})$ be a subsheaf of $X$. Then $F$ is a called a $\\tau$-open if there is a family of morphisms in $\\tau$ written $\\{X_i \\to X\\}_{i \\in I}$ such that $F$ is the image of the morphism of sheaves $\\coprod_{i \\in I} X_i \\to X$. A morphism $F \\to G$ in ${\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{A}}}^{T})$ is called is a $\\tau$-open immersion if for every $X \\in {\\text{\\bfseries\\sf{A}}}$ and every $X \\to G$ the induced morphism $F \\times_{G} X \\to X$ is a monomorphism whose image is a $\\tau$-open in $X$.\nThe category of schemes ${\\text{\\bfseries\\sf{Sch}}}({\\text{\\bfseries\\sf{A}}}, T, \\tau)$ is defined to be the full subcategory of ${\\text{\\bfseries\\sf{Sh}}}({\\text{\\bfseries\\sf{A}}}^{T})$ of sheaves $F$ such that there exists a family of $X_{i} \\in {\\text{\\bfseries\\sf{A}}}$ for $i \\in I$ and a morphism $p: \\coprod_{i \\in I} X_i \\to F$ such that $p$ is an epimorphism of sheaves and for each $i$ the morphism $X_i \\to F$ is a $\\tau$-open immersion.\n\\end{defn}\n\\begin{example}Two interesting general categories of schemes that we have in mind are ${\\text{\\bfseries\\sf{Sch}}}({\\text{\\bfseries\\sf{A}}}, T,\\tau)$ where ${\\text{\\bfseries\\sf{A}}}={\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$. First, the case where $\\tau$ of Zariski open immersions and $T$ is the Zariski pre-topology. Second, the category $\\tau$ of formal Zariski open immersions and $T$ is the formal Zariski pre-topology. \n\\end{example}\n\n\n\\begin{example}\nLet $k$ be any field and consider the category ${\\tC}= {\\text{\\bfseries\\sf{Vect}}}_{k}$. Then ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is the category of $k$-algebras. Recall that a morphism $A \\to B$ of $k$-algebras is of finite presentation in the usual sense when $B$ is finitely generated as an $A$ algebra and the ideal of relations is also finitely generated. Let us temporarily call a morphism TVfp if it satisfies the condition from Definition \\ref{defn:FinitePres}. Consider the functor\n\\[\\straightF : {\\text{\\bfseries\\sf{Set}}} \\to A\/{\\text{\\bfseries\\sf{Comm}}}({\\tC}) \n\\]\nwhich sends each set to the $A$-algebra freely generated by it. Then Lawvere's work on finitary algebraic theories and Corollary 3.13 and the remark following it in \\cite{AR} show that $A \\to B$ is TVfp if and only if there exists finite sets $S_g$ and $S_r$ and an isomorphism in $A\/{\\text{\\bfseries\\sf{Comm}}}({\\tC})$ of the form\n\\[\\colim[\\straightF S_r \\rightrightarrows \\straightF S_g ] \\to B.\n\\]\nSo a morphism $f:A \\to B$ of $k$-algebras is of finite presentation in the categorical sense if and only it is of finite presentation in terms of generators and relations. This fact also appears in the algebraic geometry literature. The implication that finite presentation in terms of generators and relations implies finite presentation in the categorical sense was shown in this case in Lemma III.8.8.2.3 of \\cite{EGA4}. For the opposite implication see \\cite{stacks-project}. \nNow \\cite{EGA4} IV.17.9.1 tells us that a morphism of schemes is a flat monomorphism, locally of finite presentation if and only if it is an open immersion. Since a morphism of affine schemes $\\spec(B) \\to \\spec(A)$ is locally of finite presentation if and only if the corresponding morphism $A \\to B$ realizes $B$ as an $A$-algebra of finite presentation we can conclude that the Zariski open immersions are precisely the standard (Zariski) open immersions in algebraic geometry. The Zariski topology in the sense of relative algebraic geometry agrees with the Zariski topology in the standard sense in the case ${\\tC}= {\\text{\\bfseries\\sf{Vect}}}_{k}$. We should remark that this is the only example in this article for which we can use Lawvere's theory and \\cite{AR} in a straightforward way. Another way to characterize the Zariski open immersion is by replacing the flat epimorphism condition with a homotopy epimorphism condition, i.e. that the natural morphism in the derived category $B\\otimes^{\\mathbb{L}}_{A}B \\to B$ is an isomorphism. It is this condition that we examine this article for the category of Banach spaces, and not the flatness condition which would give a different answer. We comment on the case of vector spaces again in Remark \\ref{rem:recover}.\n\\end{example}\n\\begin{defn}\\label{defn:PresheafZarOpIm} A morphism $f:F \\to G$ in ${\\text{\\bfseries\\sf{Pr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}}))$ is a Zariski open immersion if for every affine scheme $X$ and every morphism $X \\to G$, the induced morphism $F \\times_{G}X \\to X$ is a monomorphism of presheaves and its image agrees with the image of the map of sheaves $\\coprod_{i \\in I} X_{i } \\to X$ corresponding to a family of Zariski opens $X_{i} \\to X$. \n\\end{defn}\n\\begin{defn}\\label{defn:AffToPreFlat}\nSuppose $ F \\in {\\text{\\bfseries\\sf{Pr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}}))$ and $X \\in {\\text{\\bfseries\\sf{Aff}}}({\\tC})$. A morphism $X \\to F$ is flat if for every $Y \\in {\\text{\\bfseries\\sf{Aff}}}({\\tC})$ and every morphism $Y \\to F$ there is a Zariski open cover $\\coprod Z_{i} \\to X \\times_{F} Y$ such that the combined morphisms $Z_{i} \\to X \\times_{F} Y \\to Y$ are flat. \n\\end{defn}\n\\begin{defn}\\label{defn:PresheafFlatMor} A morphism $f:F \\to G$ in ${\\text{\\bfseries\\sf{Pr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}}))$ is flat if for every affine scheme $X$ and every morphism $X \\to G$ and every flat morphism $W \\to X \\times_{G} F$ the composition $W \\to X \\times_{G} F \\to X$ is flat.\n\\end{defn}\n\nConsider the Grothendieck site ${\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{fpqc}$. The category of simplicial objects in ${\\text{\\bfseries\\sf{Pr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}}))$ is denoted ${\\text{\\bfseries\\sf{SPr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})).$ This category comes with a (local) model structure as explained in \\cite{T}.\n\\begin{defn}\\cite{T} An object $F \\in {\\text{\\bfseries\\sf{SPr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{fpqc})$ is called a pre-Stack. An object $F \\in {\\text{\\bfseries\\sf{SPr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{fpqc})$ is called an fpqc stack if for any $X \\in {\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$ and any hypercovering $H_{\\bullet} \\to X,$ the natural morphism\n\\[F(X) \\to \\text{holim}_{[n] \\in \\Delta} F(H_n)\n\\]\nis an equivalence of simplicial sets. The category ${\\text{\\bfseries\\sf{St}}}({\\text{\\bfseries\\sf{Aff}}}({\\tC})^{fpqc}) =Ho({\\text{\\bfseries\\sf{SPr}}}({\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})^{fpqc}))$ is the category of stacks.\n\\end{defn}\n\nAn equivalent definition to the above is to define pre-stacks as functors ${\\text{\\bfseries\\sf{Aff}}}({\\tC})^{op}\\to \\infty-{\\text{\\bfseries\\sf{Gpd}}}$, where $\\infty-{\\text{\\bfseries\\sf{Gpd}}}$ is the category of infinity groupoids. This is a full subcategory of the category of $(\\infty,1)$-categories for which one could use quasi-categories. $\\infty$-groupoids in this model are Kan simplicial sets. Stacks would be pre-stacks which satisfy descent with respect to hypercovers \\cite{DHI, L1,L2, TVe2,TVe3}. Similarly we could define (pre-)stacks valued in other categories, for instance a pre-stack in categories would be a functor ${\\text{\\bfseries\\sf{Aff}}}({\\tC})^{op} \\to (\\infty,1)-{\\text{\\bfseries\\sf{Cat}}}$. Note that the category of $(1,1)$-categories embeds into the category of $(\\infty,1)$-categories. Using quasi-categories to model $(\\infty,1)$-categories, the nerve of a $1$-category is a quasi-category. We can also view dg-categories as stable quasi-categories tensored over complexes \\cite{Coh}. We will use this later to view categories of quasi-coherent $\\mathcal{O}$-modules, and $\\mathcal{D}$-modules (derived or underived) as pre-stacks valued in categories. There is an inductive definition of an $n$-algebraic stack for $n=0,1,2,\\dots$. An algebraic fpqc stack on ${\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$ is an fpqc stack on ${\\text{\\bfseries\\sf{Aff}}}(\\text{\\bfseries\\sf{C}})$ which is $n$-algebraic for some $n$.\n\n\nOne can study schemes and stacks using with the fpqc or Zariski topology we have defined above. These could be useful in the analytic context as well when ${\\tC}$ is the category of Banach spaces and one can study faithfully flat descent in this context. However, the usual G-topology that is usually studied in non-Archimedean geometry as well as the classical metric topology of complex analytic geometry are finer topologies and have smaller ``open sets\". The localizations from Definition \\ref{defn:AffLocalization} are not flat with respect to the monoidal structure on ${\\text{\\bfseries\\sf{Ban}}}_{k}$ and for this reason we need to introduce new abstractly defined topologies which will fit in well with these facts. To do so, we need to use quasi-abelian categories.\n\n\n\n\\section{Quasi-abelian categories}We review some of Schneiders' theory of quasi-abelian categories. These are special cases of Palamodov's semi-abelian categories and of pseudo-abelian categories. They also have the structure of a (Quillen) exact category in one natural way.\nThe main reference for this section is \\cite{SchneidersQA}.\n\\begin{defn}\nLet $\\mathcal{E}$ be an additive category with kernels and cokernels. A morphism $f:E\\to F$ is $\\mathcal{E}$ is called strict if the induced morphism \n\\[\\coim(f)\\to \\im(f)\n\\] is an isomorphism. \n\\end{defn}\nHere the image of $f$ is the kernel of the canonical map\n$F \\to \\coker(f)$, and the coimage of $f$ is the cokernel of the canonical map $\\ker (f)\\to E$. \n\n\\begin{defn}\\label{defn:CartCoCart}\nLet $\\mathcal{E}$ be an additive category with kernels and cokernels. We say that $\\mathcal{E}$ is quasi-abelian if it satisfies the following two conditions:\n\\begin{itemize}\n\\item In a cartesian square \n\\begin{equation*}\n\\xymatrix{ E' \\ar[r]^{f'} \\ar[d] & F' \\ar[d] \\\\\nE \\ar[r]_{f}& F}\n\\end{equation*}\n\nIf $f$ is a strict epimorphism then $f'$ is a strict epimorphism.\n\n\\item In a co-cartesian square\n\\begin{equation*}\n\\xymatrix{ E \\ar[r]^{f} \\ar[d] & F \\ar[d] \\\\\nE' \\ar[r]_{f'}& F'}\n\\end{equation*}\nIf $f$ is a strict monomorphism then $f'$ is a strict monomorphism.\n\n\\end{itemize}\n\\end{defn}\n\\begin{rem}\\label{rem:LeftRightInverse} Any morphism in a quasi-abelian category with a right inverse is a strict epimorphism. Any morphism in a quasi-abelian category with a left inverse is a strict monomorphism. \\end{rem}\n\\begin{defn}\nLet ${\\text{\\bfseries\\sf{E}}}$ be a quasi-abelian category. \nLet $\\xymatrix{ E'\\ar[r]^{e'} & E\\ar[r]^{e''} & E''}$ be a sequence of maps such that $e''\\circ e'=0$. We call such a sequence strictly exact (resp. strictly coexact) if $e'$ (resp. $e''$) is strict and the canonical map \n$\\im(e')\\to \\ker(e'')$ is an isomorphism. A complex \n$E_1\\to \\cdots \\to E_n$ is strictly exact (resp. strictly coexact) if each subsequence $E_{i-1}\\to E_i\\to E_{i+1}$ is strictly exact (resp. strictly coexact).\n\\end{defn}\n\n\\begin{rem}\\label{rem:threetermstrict}\nNote that the sequence \\begin{equation}\\label{equation:threetermstrict}\\xymatrix{ 0\\ar[r] & E\\ar[r]^{u}& F\\ar[r]^{v}& G\\ar[r]& 0}\\end{equation} is strictly exact if and only if $u$ is the kernel of $v$ and $v$ is the cokernel of $u$. Any strict monomorphism or strict epimorphism can be completed to a strictly exact sequence in the form of Equation \\ref{equation:threetermstrict}. \nThis implies that such a sequence is strictly exact if and only if it is strictly coexact.\n\\end{rem}Let ${\\text{\\bfseries\\sf{E}}}$ be a closed symmetric monoidal quasi-abelian category with all finite limits and colimits. Then an object is flat if and only if tensoring with it preserves strict short exact sequences.\n\n\\begin{defn} Call a sequence $\\xymatrix{ E'\\ar[r]^{e'} & E\\ar[r]^{e''} & E''}$ exact (resp. coexact) if the canonical map \n$\\im(e')\\to \\ker(e'')$ is an isomorphism. A sequence \n$E_1\\to \\cdots \\to E_n$ is exact (resp. coexact) if each subsequence $E_{i-1}\\to E_i\\to E_{i+1}$ is exact (resp. coexact).\n\\end{defn}\n\\begin{rem}\\label{rem:threeterm}\nNote that the sequence \\begin{equation}\\label{equation:threeterm}\\xymatrix{ 0\\ar[r] & E\\ar[r]^{u}& F\\ar[r]^{v}& G\\ar[r]& 0}\\end{equation} is exact if and only if $\\ker(u)=0$, $\\im(v)=G$ and $\\im(u) \\to \\ker(v)$ is an isomorphism. Any monomorphism or epimorphism can be completed to a exact sequence in the form of Equation \\ref{equation:threeterm}. \n\\end{rem}\n\nThe following is remark $1.1.11$ in \\cite{SchneidersQA}:\n\\begin{thm}\nLet ${\\text{\\bfseries\\sf{E}}}$ be a quasi-abelian category. The class of strictly exact short exact sequences endows ${\\text{\\bfseries\\sf{E}}}$ with the structure of an exact category.\n\\end{thm}\n\n\\begin{defn}\nLet ${\\text{\\bfseries\\sf{E}}}$ be a quasi-abelian category. Let $\\straightK({\\text{\\bfseries\\sf{E}}})$ be its homotopy category. The derived category of ${\\text{\\bfseries\\sf{E}}}$ is $\\straightD({\\text{\\bfseries\\sf{E}}})=\\straightK({\\text{\\bfseries\\sf{E}}})\/\\straightN({\\text{\\bfseries\\sf{E}}})$ where $\\straightN({\\text{\\bfseries\\sf{E}}})$ is the full subcategory of strictly exact sequences.\n\\end{defn}\n\n\\begin{defn}\nLet ${\\text{\\bfseries\\sf{E}}}$ be a quasi-abelian category. Let $\\straightK({\\text{\\bfseries\\sf{E}}})$ be its homotopy category. A morphism in \n$\\straightK({\\text{\\bfseries\\sf{E}}})$ is called a strict quasi-isomorphism if its mapping cone is strictly exact. \n\\end{defn}\n\nThe following is 1.2.17, 1.2.19, 1.2.20, 1.2.27 and 1.2.31 in \\cite{SchneidersQA}: \n\\begin{thm}\nLet ${\\text{\\bfseries\\sf{E}}}$ be a quasi-abelian category.\n\\begin{enumerate}\n\\item $\\straightD({\\text{\\bfseries\\sf{E}}})$ has a canonical t-structure (the left t-structure). A complex $E$ belongs to $D^{\\leq 0}$ if and only if it is strictly exact in strictly positive degrees. $E$ belongs to $D^{\\geq 0}$ if and only if it is strictly exact in strictly negative degrees.\n\\item The heart of this t-structure $\\LH({\\text{\\bfseries\\sf{E}}})$, is equivalent to the localization of the full subcategory of $\\straightK({\\text{\\bfseries\\sf{E}}})$ consisting of complexes E of the form \n\\begin{equation}\n\\xymatrix{ 0\\ar[r] & E\\ar[r]^{u}& F\\ar[r]& 0}\n\\end{equation}\nwhere $u$ is a monomorphism and $F$ is in degree $0$, by the multiplicative system formed by morphisms which are both cartesian and cocartesian. \n\\item There is a canonical fully faithful functor $\\sI:{\\text{\\bfseries\\sf{E}}}\\to LE({\\text{\\bfseries\\sf{E}}})$. A sequence $E'\\to E\\to E''$ is strictly exact in ${\\text{\\bfseries\\sf{E}}}$ if and only if $\\sI(E')\\to \\sI(E)\\to \\sI(E'')$ is exact in $\\LH({\\text{\\bfseries\\sf{E}}})$.\n\\item The functor $I$ induces an equivalence between $\\straightD({\\text{\\bfseries\\sf{E}}})$ and $\\straightD(\\LH({\\text{\\bfseries\\sf{E}}}))$. This equivalence sends the (left) t-structure on $\\straightD({\\text{\\bfseries\\sf{E}}})$ to the standard t-structure on $\\straightD(\\LH({\\text{\\bfseries\\sf{E}}}))$.\n\\end{enumerate}\n\\end{thm}\n\n\\begin{rem}\nThe embedding $\\sI:{\\text{\\bfseries\\sf{E}}}\\to \\LH({\\text{\\bfseries\\sf{E}}})$ is universal \nin the sense that induces an equivalence for any abelian category $\\mathcal{F}$ between left strictly exact functors from ${\\text{\\bfseries\\sf{E}}}$ to $\\mathcal{F}$ and left exact functors \nfrom $\\LH({\\text{\\bfseries\\sf{E}}})$ to $\\mathcal{F}$. In this sense $\\LH({\\text{\\bfseries\\sf{E}}})$ is the (left) abelian envelope of ${\\text{\\bfseries\\sf{E}}}$. See 1.2.33 in \\cite{SchneidersQA}.\n\\end{rem}\n\n\\begin{defn}\\label{defn:Projective}\nLet ${\\text{\\bfseries\\sf{E}}}$ be an additive category with kernels and cokernels. An object $I$ is called injective (resp. strongly injective) if the functor $E\\mapsto \\Hom(E,I)$ is exact (resp. strongly exact), i.e. for any strict (resp. arbitrary) monomorphism $u:E\\to F$, the induced map $\\Hom(F,I)\\to \\Hom(E,I)$ is surjective. Dually, $P$ is called projective (resp. strongly projective) if the functor $E\\mapsto \\Hom(P,E)$ is exact (resp. strongly exact), i.e. for any strict (resp. arbitrary) epimorphism $u:E\\to F$, the associated map $\\Hom(P,E)\\to \\Hom(P,F)$ is surjective. \n\\end{defn}\n\n\\begin{defn}\\label{defn:enough}\nA quasi-abelian category ${\\text{\\bfseries\\sf{E}}}$ has enough projectives if for any object $E$ there is a strict epimorphism $P\\to E$ where $P$ is projective. A quasi-abelian category ${\\text{\\bfseries\\sf{E}}}$ has enough injectives if for any object $E$ there is a strict monomorphism $E \\to I$ where $I$ is injective.\n\\end{defn}\n\nThe following is 1.3.24 in \\cite{SchneidersQA}:\n\\begin{lem}\nLet ${\\text{\\bfseries\\sf{E}}}$ be a quasi-abelian category. \n\\begin{enumerate}\n\\item An object $P$ of ${\\text{\\bfseries\\sf{E}}}$ is projective if and only if $\\sI(P)$ is projective in $\\LH({\\text{\\bfseries\\sf{E}}})$.\n\\item ${\\text{\\bfseries\\sf{E}}}$ has enough projectives if and only if $\\LH({\\text{\\bfseries\\sf{E}}})$ has enough projectives. In this case an object of $\\LH({\\text{\\bfseries\\sf{E}}})$ is projective if it is isomorphic to $\\sI(P)$ where $P$ is projective in ${\\text{\\bfseries\\sf{E}}}$.\n\\end{enumerate}\n\\end{lem}\n\nThe following is 1.3.22 in \\cite{SchneidersQA}:\n\\begin{thm}\nLet ${\\text{\\bfseries\\sf{E}}}$ be a quasi-abelian category with enough projectives (resp. injectives). Let $\\text{\\bfseries\\sf{P}}$ be the full additive subcategory of projective objects (resp. $\\text{\\bfseries\\sf{I}}$ the category of injective objects). The canonical functor $\\sK^-(\\text{\\bfseries\\sf{P}})\\to \\sD^-({\\text{\\bfseries\\sf{E}}})$ (resp. $\\sK^+(\\text{\\bfseries\\sf{I}})\\to \\sD^+({\\text{\\bfseries\\sf{E}}})$) is an equivalence.\n\\end{thm}\n\n\n\n\n\\subsection{Closed symmetric monoidal quasi-abelian categories.}\n\nThe following is 1.5.1 in \\cite{SchneidersQA}:\n\\begin{prop}\\label{prop:StrictIFF}\nSuppose that ${\\tC}$ is a closed symmetric monoidal quasi-abelian category with all finite limits and colimits. Suppose that $A \\in {\\text{\\bfseries\\sf{Comm}}}({\\tC})$. The category ${\\text{\\bfseries\\sf{Mod}}}(A)$ \nis quasi-abelian and the forgetful functor ${\\text{\\bfseries\\sf{Mod}}}(A)\\to {\\tC}$ preserves limits and colimits. A morphism in ${\\text{\\bfseries\\sf{Mod}}}(A)$ is strict if and only if it is strict in ${\\tC}$. \n\\end{prop}\n\n\n\\begin{lem}Suppose that ${\\tC}$ is a closed symmetric monoidal quasi-abelian category with all finite limits and colimits. Using Remark \\ref{rem:LeftRightInverse} we see that for any $V \\in {\\text{\\bfseries\\sf{Mod}}}(A)$ the canonical morphism\n\\[V \\overline{\\otimes} A \\to V\n\\]\nis a strict epimorphism and the canonical morphism\n\\[V \\to \\underline{\\Hom}(A,V)\n\\]\nis a strict monomorphism.\n\\end{lem}\n\n\n\\begin{defn}\\label{defn:finiteAmodAbstract}\nSuppose that ${\\tC}$ is a closed symmetric monoidal quasi-abelian category with all finite limits and colimits. An object $V$ is called finite if there is a strict epimorphism $\\coprod_{i=1}^{n} \\text{id}_{{\\tC}} \\to V$ in ${\\tC}$ for some finite non-negative integer $n$. In the case that ${\\tC}={\\text{\\bfseries\\sf{Mod}}}(A)$ for $A$ a commutative monoid in a closed symmetric monoidal quasi-abelian category, we denote the full subcategory of finite objects by ${\\text{\\bfseries\\sf{Mod}}}^{f}(A)$.\n\\end{defn}\n\n\\begin{lem}\\label{lem:FreeProjCofreeInj} Suppose that $0 \\to L \\to M \\to N \\to 0$ is a strictly exact sequence in ${\\text{\\bfseries\\sf{Mod}}}(A)$ and $F =A\\overline{\\otimes} P \\in {\\text{\\bfseries\\sf{Mod}}}(A)$ is free and $C=\\underline{\\Hom}(A,I) \\in {\\text{\\bfseries\\sf{Mod}}}(A)$ is cofree. Then the sequences \n\\[0 \\to \\Hom_{A}(F,L) \\to \\Hom_{A}(F,M)\\to \\Hom_{A}(F,N) \\to 0\n\\]\nand\n\\[0 \\to \\Hom_{A}(N,C) \\to \\Hom_{A}(M,C)\\to \\Hom_{A}(L,C) \\to 0\n\\]\nare exact. If the sequence $0 \\to L \\to M \\to N \\to 0$ is only exact and $F$ is strictly free and $C$ is strictly cofree we can make the same conclusion.\n\\end{lem}\n{\\bf Proof.}\nUsing Lemma \\ref{lem:MeyerProperties} these sequences are isomorphic to the sequences\n\\[0 \\to \\Hom(P,L) \\to \\Hom(P,M)\\to \\Hom(P,N) \\to 0\n\\]\nand \n\\[0 \\to \\Hom(N,I) \\to \\Hom(M,I)\\to \\Hom(L,I) \\to 0\n\\]\nwhich are exact by definition of projectivity and injectivity (or the strict versions).\n\\hfill $\\Box$\n\\begin{lem}\\label{lem:MaintainProjInj}If $P$ is projective in ${\\tC}$ then $P\\overline{\\otimes} A$ is projective in ${\\text{\\bfseries\\sf{Mod}}}(A)$. Similarly, if $I$ is injective in ${\\tC}$ then $\\underline{\\Hom}(A,I)$ is injective in ${\\text{\\bfseries\\sf{Mod}}}(A)$.\n\\end{lem}\n{\\bf Proof.}\nBoth of these facts are immediately implied by Lemma \\ref{lem:FreeProjCofreeInj} together with Remark \\ref{rem:threetermstrict}.\n\\hfill $\\Box$\n\\begin{lem} \\label{lem:MonoEpiPreserved} The functor $E \\mapsto E\\overline{\\otimes} A$ takes epimorphisms in ${\\tC}$ to epimorphisms in ${\\text{\\bfseries\\sf{Mod}}}(A)$. The functor $E \\mapsto \\underline{\\Hom}(A,E)$ takes monomorphisms in ${\\tC}$ to monomorphisms in ${\\text{\\bfseries\\sf{Mod}}}(A)$.\n\\end{lem}\n{\\bf Proof.}\nThis follows directly from the definitions and Lemma \\ref{lem:MeyerProperties}.\n\\hfill $\\Box$\n\n\\begin{defn}\\label{defn:KerFlat} For $A \\in {\\text{\\bfseries\\sf{Comm}}}({\\tC})$ a module $M$ in ${\\text{\\bfseries\\sf{Mod}}}(A)$ is called kernel flat if for any morphism $f:E \\to F$ in ${\\text{\\bfseries\\sf{Mod}}}(A)$ the natural morphism \n\\begin{equation}\\label{equation:want4flatAbstr}B\\overline{\\otimes} \\ker(f) \\to \\ker(f_{M})\n\\end{equation}\nis an isomorphism where $f_{M}$ is defined as $\\text{id}_{M} \\overline{\\otimes}_{A}f:M \\overline{\\otimes}_{A}E \\to M \\overline{\\otimes}_{A}F.$ A morphism $A \\to B$ in ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is called kernel flat if it makes $B$ kernel flat over $A$.\n\\end{defn}\n\n\\begin{lem}\\label{lem:FlatkerFlat}Suppose that ${\\tC}$ is a closed symmetric monoidal quasi-abelian category. An object $V\\in {\\tC}$ is kernel flat if and only it is flat (see Definition \\ref{defn:AbstractFlat}). Therefore, a morphism of algebras $p: A \\to B$ is kernel flat (see Definition \\ref{defn:KerFlat}) if and only if it is flat (Definition \\ref{defn:AbstractFlat}) in ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$. \n\n\\end{lem}\n\n{\\bf Proof.}\nFirst of all if $p$ is flat it is clearly kernel flat since a kernel is a type of limit. In the other direction suppose that $p$ is kernel flat. It means that tensoring with $B$ commutes with kernels. Note that every limit over a finite diagram can be written as a combination of finite products and kernels. Finite products are isomorphic to finite coproducts and the functor given by tensoring with $B$ commutes with coproducts and hence it commutes with finite products. Therefore, tensoring with $B$ commutes with finite limits and hence $p$ is flat. \n\n\\ \\hfill $\\Box$\n\\begin{lem}\\label{lem:HelpWithEnough} Suppose that ${\\tC}$ is a closed symmetric monoidal quasi-abelian category with all finite limits and colimits. Suppose that $A \\in {\\text{\\bfseries\\sf{Comm}}}({\\tC})$. If the category ${\\tC}$ has enough projectives \nthen the category ${\\text{\\bfseries\\sf{Mod}}} (A)$ has enough projectives. If the category ${\\tC}$ has enough injectives \nthen the category ${\\text{\\bfseries\\sf{Mod}}}(A)$ has enough injectives. \n\\end{lem}\n{\\bf Proof.} Suppose that ${\\tC}$ has enough projectives. Suppose that $V \\in {\\text{\\bfseries\\sf{Mod}}}(A)$. Choose a strict epimorphism in ${\\tC}$ of the form $P \\to V$ where $P$ is projective in ${\\tC}$. Lemma \\ref{lem:MaintainProjInj} implies that $P\\overline{\\otimes} A$ is projective in ${\\text{\\bfseries\\sf{Mod}}}(A).$ Consider the morphism $P\\overline{\\otimes} A \\to V.$ We need to show it is a strict epimorphism in ${\\text{\\bfseries\\sf{Mod}}}(A).$ It factorizes as \n\\begin{equation}\nP \\overline{\\otimes} A \\to V \\overline{\\otimes} A \\to V.\n\\end{equation}\nThe second morphism is a strict epimorphism because it admits a right inverse. The arrow $P \\overline{\\otimes} A \\to V \\overline{\\otimes} A$ is an epimorphism by Lemma \\ref{lem:MonoEpiPreserved} and in fact a strict epimorphism because the monoidal product with $A$ is a left adjoint functor and preserves cokernels. Therefore ${\\text{\\bfseries\\sf{Mod}}}(A)$ has enough projectives.\nSuppose that ${\\tC}$ has enough injectives. Choose a strict monomorphism in ${\\tC}$ of the form $V \\to I$ where $I$ is injective in ${\\tC}$. Lemma \\ref{lem:MaintainProjInj} implies that $\\underline{\\Hom}(A,I)$ is injective in ${\\text{\\bfseries\\sf{Mod}}}(A).$ Consider the morphism $V \\to \\underline{\\Hom}(A,I).$ We need to show that it is a strict monomorphism in ${\\text{\\bfseries\\sf{Mod}}}(A).$ It factorizes as \\begin{equation}\nV \\to \\underline{\\Hom}(A,V) \\to \\underline{\\Hom}(A,I).\n\\end{equation}\nNotice that here, we are considering $\\underline{\\Hom}(A,V)$ and $\\underline{\\Hom}(A,I)$ as elements of \n${\\text{\\bfseries\\sf{Mod}}}(A)$ using the action of $A$ on itself. The first arrow is a strict monomorphism because it admits a left inverse. Using Lemma \\ref{lem:MonoEpiPreserved}, $\\underline{\\Hom}(A,V) \\to \\underline{\\Hom}(A,I)$ is a monomorphism in ${\\text{\\bfseries\\sf{Mod}}}(A)$ and in fact a strict monomorphism because the internal Hom from $A$ is a right adjoint functor and preserves kernels. Therefore ${\\text{\\bfseries\\sf{Mod}}}(A)$ has enough injectives.\n\\hfill $\\Box$\n\\begin{lem}\\label{lem:BaseChangeInj}Suppose that $\\sF: {\\tC} \\to \\text{\\bfseries\\sf{D}}$ is a functor which has a right adjoint $\\sR$ and which \npreserves strict monomorphisms (preserves monomorphisms). Then an injective (strongly injective) in $\\text{\\bfseries\\sf{D}}$ is an injective (strongly injective) when considered in ${\\tC}$ via $\\sR$ .\n\\end{lem}\n{\\bf Proof.}\n\nSuppose that $I \\in \\text{\\bfseries\\sf{D}}$ is injective (strongly injective). Then consider a strict monomorphism (monomorphism) $E\\to F$ in ${\\tC}.$ Then $\\sF(E) \\to \\sF(F)$ is a strict monomorphism (monomorphism) in ${\\tC}.$ We have a commutative diagram\n\\[\\xymatrix{{\\tC}(F,\\sR (I)) \\ar[r]& {\\tC} (E,\\sR (I)) \\\\\n\\text{\\bfseries\\sf{D}}(\\sF (F),I) \\ar[r] \\ar[u]& \\text{\\bfseries\\sf{D}} (\\sF (E),I). \\ar[u]\n}\n\\]\nBecause the upwards arrows are isomorphisms and the lower horizontal arrow is surjective, the upper horizontal arrow is surjective as well. Therefore, $I$ is injective (strongly injective) when considered as an object of ${\\tC}$. \n\\hfill $\\Box$\n\n\\begin{lem}\\label{lem:BaseChangeProj}Suppose that $\\sG: {\\tC} \\to \\text{\\bfseries\\sf{D}}$ is a functor which has a left adjoint $\\sL$ and which preserves strict epimorphisms (preserves epimorphisms). Then a projective (strongly projective) in $\\text{\\bfseries\\sf{D}}$ is projective (strongly projective) when considered in ${\\tC}$.\n\\end{lem}\n{\\bf Proof.}\nSuppose that $P \\in \\text{\\bfseries\\sf{D}}$ is projective (strongly projective). Consider a strict epimorphism (epimorphism) $E\\to F$ in $\\text{\\bfseries\\sf{D}}(A).$ Then by $\\sG(E) \\to \\sG(F)$ is a strict epimorphism (epimorphism) in $\\text{\\bfseries\\sf{D}}.$ We have a commutative diagram\n\\[\\xymatrix{{\\tC}(\\sL (P),E) \\ar[r]& {\\tC} (\\sL (P),F)\\\\\n\\text{\\bfseries\\sf{D}}(P,\\sG (E))\\ar[u] \\ar[r]& \\text{\\bfseries\\sf{D}}(P,\\sG (F)) \\ar[u].\n}\n\\]\nBecause the upwards arrows are isomorphisms and the lower horizontal arrow is surjective, the upper horizontal arrow is surjective as well. Therefore, $P$ is projective (strongly projective) when considered in ${\\tC}$. \n\\ \\hfill $\\Box$\n\nLet ${\\text{\\bfseries\\sf{E}}}$ be a closed symmetric monoidal quasi-abelian category and let $A\\in {\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{E}}})$. \nThe following is contained in 2.1.18 in \\cite{SchneidersQA}: If $P$ is projective in ${\\text{\\bfseries\\sf{E}}}$ then $A\\overline{\\otimes} P$ is projective in ${\\text{\\bfseries\\sf{Mod}}}(A)$.\n\n\n\n \n\\subsection{Derived Functors}\nLet $\\sF: {\\tC} \\to {\\tD}$ be an additive functor betwen quasi-abelian categories ${\\tC}$ and ${\\tD}$. Schneiders gave the following definitions in 1.3.2 of \\cite{SchneidersQA}\n\\begin{defn} A full additive subcategory ${\\tP}$ of ${\\tC}$ is called $\\sF$-projective if:\n\\begin{enumerate}\n\\item for any object $V$ of ${\\tC}$ there is an object $P$ of ${\\tP}$ and a strict epimorphism $P\\to V$\n\\item in any strictly exact sequence \n\\[0 \\to V' \\to V \\to V'' \\to 0\n\\]\nof ${\\tC}$ where $V$ and $V''$ are objects of ${\\tP}$, $V'$ is as well\n\\item for any strictly exact sequence \n\\[0 \\to V' \\to V \\to V'' \\to 0\n\\]\nof ${\\tC}$ where $V, V'$ and $V''$ are objects of ${\\tP}$, the sequence \n\\[0 \\to \\sF(V') \\to \\sF(V) \\to \\sF(V'') \\to 0\n\\]\nis strictly exact in ${\\tD}.$\n\\end{enumerate}\n A full additive subcategory ${\\tI}$ of ${\\tC}$ is called $\\sF$-injective if:\n\\begin{enumerate}\n\\item for any object $V$ of ${\\tC}$ there is an object $I$ of ${\\tI}$ and a strict monomorphism $V\\to I$\n\\item in any strictly exact sequence \n\\[0 \\to V' \\to V \\to V'' \\to 0\n\\]\nof ${\\tC}$ where $V$ and $V''$ are objects of ${\\tI}$, $V'$ is as well\n\\item for any strictly exact sequence \n\\[0 \\to V' \\to V \\to V'' \\to 0\n\\]\nof ${\\tC}$ where $V, V'$ and $V''$ are objects of ${\\tI}$, the sequence \n\\[0 \\to \\sF(V') \\to \\sF(V) \\to \\sF(V'') \\to 0\n\\]\nis strictly exact in ${\\tD}.$\n\\end{enumerate}\n\\end{defn}\nSchneiders also includes the following (Lemma 1.3.3 \\cite{SchneidersQA})\n\\begin{lem}\nLet ${\\tC}$ be a quasi-abelian category and let ${\\tP}$ be a subset of the objects of ${\\tC}$. Assume that for any object $V$ of ${\\tC}$ there is a strict epimorphism $P \\to V$ with $P \\in {\\tP}$. Then for each object $V$ of $C^{-}({\\tC})$ there is a quasi-isomorphism $u:P \\to V$ with $P$ in $C^{-}({\\tP})$ and such that each $u^{k}:P^{k} \\to V^{k}$ is a strict epimorphism.\n\\end{lem}\n\nFrom this we get (proposition 1.3.5 \\cite{SchneidersQA}):\n\\begin{prop}\\label{lem:derivable}\nLet $\\sF: {\\tC} \\to {\\tD}$ be an additive functor between quasi-abelian categories ${\\tC}$ and ${\\tD}$.\n\\begin{enumerate}\n\\item Assume that ${\\tC}$ has an $\\sF$-projective subcategory. Then $\\sF$ has a left derived functor $L\\sF:D^-({\\tC})\\to D^-({\\tD})$.\n\\item Assume that ${\\tC}$ has an $\\sF$-injective subcategory. Then $\\sF$ has a right derived functor $R\\sF:D^+({\\tC})\\to D^+({\\tD})$.\n\\end{enumerate}\n\\end{prop}\n\\begin{defn}In the situations of Lemma \\ref{lem:derivable}, $\\sF$ is called explicitly left derivable or explicitly right derivable.\n\\end{defn}\nHere, derived functors are defined as usual by their universal property.\n\n\\begin{rem}\\label{rem:sse2sse}Note that if $\\sF$ is exact (sends strict short exact sequences to strict short exact sequences) then the full subcategory ${\\tC}$ itself is an $\\sF$-projective (and injective) subcategory. Hence exact functors are always derivable. \n\\end{rem}\n\n\n\nAs in the abelian case, projective and injectives form $\\sF$-projective and $\\sF$-injective subcategories (remark 1.3.21 \\cite{SchneidersQA}):\n\\begin{prop}\\label{prop:ProjFProj}\nLet $\\sF: {\\tC} \\to {\\tD}$ be an additive functor between quasi-abelian categories ${\\tC}$ and ${\\tD}$. \n\\begin{enumerate}\n\\item Assume that ${\\tC}$ has enough projectives. Then the full subcategory of projective objects is a $\\sF$-projective subcategory and therefore can be used to explicitly left derive the functor $\\sF.$\n\\item Assume that ${\\tC}$ has enough injectives. Then the full subcategory of injective objects is a $\\sF$-injective subcategory and therefore can be used to explicitly right derive the functor $\\sF.$\n\\end{enumerate}\n\\end{prop}\nWe also have the following (remark 1.3.7 \\cite{SchneidersQA}):\n\\begin{lem}\\label{lem:acyclic}\nLet $\\sF: {\\tC} \\to {\\tD}$ be an additive functor between quasi-abelian categories ${\\tC}$ and ${\\tD}$. Assume that $\\sF$ has a right derived functor $R\\sF:D^+({\\tC})\\to D^+({\\tD})$. Call an object $I$ $\\sF$-acyclic if $R\\sF(I)\\cong \\sF(I)$. Assume that for any object \n$A$, there is an $\\sF$-acyclic object $I$ and a monomorphism \n$A\\to I$. Then the $\\sF$-acyclic objects form a $\\sF$-injective subcategory. Assume that $\\sF$ has a left derived functor $L\\sF:D^-({\\tC})\\to D^-({\\tD})$. Call an object $P$ $\\sF$-acyclic if $L\\sF(P)\\cong \\sF(P)$. Assume that for any object \n$A$, there is an $\\sF$-acyclic object $P$ and a epimorphism \n$P \\to A$. Then the $\\sF$-acyclic objects form a $\\sF$-projective subcategory. \n\\end{lem}\n\\begin{defn}Let ${\\tC}$ be a closed symmetric monoidal quasi-abelian category with monoidal structure $\\overline{\\otimes}$. An object $V$ of ${\\tC}$ is called $\\overline{\\otimes}$-acyclic if $V$ is $\\sF$-acyclic for all of the functors $\\sF:{\\tC} \\to {\\tC}$ given by $U \\mapsto U \\overline{\\otimes} W$ for any object $W$ in ${\\tC}$.\n\\end{defn}\n\\subsection{Topologies based on homological algebra}\nUsing the homological algebra in this section, we now introduce some more classes of morphisms and Grothendieck topologies on a closed symmetric monoidal quasi-abelian category $\\text{\\bfseries\\sf{C}}$ with all finite limits and colimits.\n\\begin{lem}For any morphism $p:\\spec(B)\\to \\spec(A)$ in ${\\text{\\bfseries\\sf{Aff}}}({\\tC}),$ the induced morphism $p_{*}:{\\text{\\bfseries\\sf{Mod}}}(B)\\to {\\text{\\bfseries\\sf{Mod}}}(A)$ is derivable to a functor $\\sD^{-}(B) \\to \\sD^{-}(A).$ \n\\end{lem}\n{\\bf Proof.} This functor sends strict exact sequences to strict exact sequences so this follows from Remark \\ref{rem:sse2sse}.\n\\ \\hfill $\\Box$\n\\begin{defn}\\label{defn:homotopyEpi}A morphism $p:\\spec(B)\\to \\spec(A)$ in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ is called a homotopy monomorphism in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ if the induced morphism $p_{*}:\\sD^{-}(B) \\to \\sD^{-}(A)$ is fully faithful.\n\\end{defn}\nNotice that by considering $i=0$ in Definition \\ref{defn:homotopyEpi} we see that a homotopy epimorphism in ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is in particular an epimorphism in ${\\text{\\bfseries\\sf{Comm}}}({\\tC}).$\n\n\\begin{lem}\\label{lem:ComposHom}The composition of homotopy monomorphisms in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ is a homotopy monomorphism in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$.\n\\end{lem}\n{\\bf Proof.}\nThis follows from the fact that the composition of fully faithful functors is fully faithful.\n\\ \\hfill $\\Box$\n\\begin{lem}\n\nAssume that $p:\\spec(B)\\to \\spec(A)$ in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ and that the functor $p^{*}:{\\text{\\bfseries\\sf{Mod}}}(A) \\to {\\text{\\bfseries\\sf{Mod}}}(B)$ given by tensoring with $B$ over $A$ is explicitly left derivable to a functor $\\mathbb{L}p^{*}:\\sD^{-}(A)\\to \\sD^{-}(B)$. Then $p$ is homotopy monomorphism if and only if the natural morphism of functors $\\mathbb{L}p^{*}p_{*}\\to id_{\\sD^{-}(B)}$ is an isomorphism.\n\n\\end{lem}\n{\\bf Proof.}\nWe have natural isomorphisms for any objects $M,N \\in \\sD^{-}(B)$ \n\\[\\Hom_{\\sD^{-}(B)}(\\mathbb{L}p^{*}p_{*}M,N) \\cong \\Hom_{\\sD^{-}(A)}(p_{*}M,p_{*}N) .\n\\]\nTherefore, if $\\mathbb{L}p^{*}p_{*}\\to id_{\\sD^{-}(B)}$ is a isomorphism then $p$ is a homotopy epimorphism. The converse follows from a simple application of the Yoneda lemma.\n\\ \\hfill $\\Box$\n\\begin{lem}\\label{lem:HomotopyMon}\n Assume that $p:\\spec(B)\\to \\spec(A)$ is a morphism in ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ and that the functor ${\\text{\\bfseries\\sf{Mod}}}(A) \\to {\\text{\\bfseries\\sf{Mod}}}(B)$ given by tensoring with $B$ over $A$ is explicilty left derivable to a functor $\\sD^{-}(A)\\to \\sD^{-}(B)$. Then $p$ is homotopy monomorphism if and only if $B\\overline{\\otimes}^{\\mathbb{L}}_{A}B\\cong B$.\n\\end{lem}\n{\\bf Proof.} \nFor any object $M$ of $\\sD^{-}(B)$ we have \n\\begin{equation*}\nM\\overline{\\otimes}^{\\mathbb{L}}_{A}B\\cong M\\overline{\\otimes}^{\\mathbb{L}}_{B}(B\\overline{\\otimes}^{\\mathbb{L}}_{A}B).\n\\end{equation*}\n\nHence $\\mathbb{L}p^{*}p_{*}\\to id_{\\sD^{-}(B)}$ is an isomorphism if and only if we have natural isomorphisms $M\\overline{\\otimes}^{\\mathbb{L}}_{A}B\\cong M$ for any $M \\in \\sD^{-}(B)$ which happens if and only if $B\\overline{\\otimes}^{\\mathbb{L}}_{A}B\\cong B$.\n\\ \\hfill $\\Box$\n\n\n\\begin{defn}\\label{defn:fhTVZ} Let ${\\tC}$ be a closed, symmetric monoidal quasi-abelian category with enough projectives.\n The morphism $\\spec(B) \\to \\spec(A)$ of ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ is called a homotopy formal Zariski open immersion if the corresponding morphism $A \\to B$ in ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is a homotopy epimorphism.\n\\end{defn}\n\\begin{defn}\\label{defn:hTVZ} Let ${\\tC}$ be a closed, symmetric monoidal quasi-abelian category with enough projectives.\n The morphism $\\spec(B) \\to \\spec(A)$ of ${\\text{\\bfseries\\sf{Aff}}}({\\tC})$ is called a homotopy Zariski open immersion if the corresponding morphism $A \\to B$ in ${\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is a homotopy epimorphism of finite presentation.\n\\end{defn}\n\n\\begin{defn}\\label{defn:Amitsur}The Amitsur complex of a morphism $f:A \\to B$ in $ {\\text{\\bfseries\\sf{Comm}}}({\\tC})$ is the complex $\\mathscr{A}(f)$ given by \n\\[0 \\to A \\to B \\to B\\overline{\\otimes}_{A}B \\to B\\overline{\\otimes}_{A}B\\overline{\\otimes}_{A}B \\to \\cdots\n\\]\nwhere the morphism $B^{\\overline{\\otimes}_{A}^{m}}\\to B^{\\overline{\\otimes}_{A}^{m+1}}$ is defined by \n\\[d(b_{1} \\overline{\\otimes} b_{2} \\overline{\\otimes} \\cdots \\overline{\\otimes} b_{m})= \\sum_{i=1}^{m+1}(-1)^{i}b_{1} \\overline{\\otimes} \\cdots \\overline{\\otimes} b_{i-1}\\otimes 1 \\overline{\\otimes} b_{i} \\overline{\\otimes} \\cdots \\overline{\\otimes} b_{m}.\n\\]\nIn the case that we can chose a decomposition $f= \\prod_{i=1}^{n} f_{i}:A \\to \\prod_{i=1}^{n} B_{i}=B$ there is a strictly included subcomplex $\\mathscr{A}^{a}(f)\\to \\mathscr{A}(f)$ where $\\mathscr{A}^{a}(f)$ is given by the finite complex \n\\[0 \\to A \\to \\prod_{1 \\leq i_1 \\leq n} B_{i} \\to \\prod_{1\\leq i_10$ and $(f_1, \\dots, f_m,g)=1.$\nNotice that following Proposition 1 of 7.2.4 of \\cite{BGR} $|g|$ cannot be arbitrarily small for $|\\;| \\in \\mathcal{M}(\\mathcal{A}_{V})$ and in fact we can realize the rational localization as an Laurent localization of a Weierstrass localization. That is, there is an $\\epsilon>0$ such that\n\\[\\mathcal{A}_{V}\n\\cong \\left(\\mathcal{A}\\{ \\frac{S}{\\epsilon^{-1}}\\}\/(gS-1)\\right)\\{\\frac{T_1}{r_1}, \\dots, \\frac{T_m}{r_m} \\}\/( T_1 - \\frac{f'_1}{g'}, \\dots, T_m -\\frac{f'_m}{g'})\n\\]\nwhere the $f'_i$ and $g'$ are the image of $f_i$ and $g$ in $\\mathcal{A}\\{\\epsilon S\\}\/(gS-1).$ This corresponds to the fact that we can choose $\\epsilon >0$ such that the conditions $|f_{i}|\\leq r_{i}|g|$ and $(f_1, \\dots, f_m,g)=1$ are equivalent to $|g|>\\epsilon$ and $|\\frac{f_{i}}{g}|\\leq r_{i}$. Therefore Lemma \\ref{lem:easyLW} implies that the natural morphism\n\\[\\mathcal{A}_{V}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{B} \\to \\mathcal{A}_{V}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B} \\] \nis an isomorphism in $\\sD^{-}(\\mathcal{A})$. \nNow if $k$ is trivially valued, consider, using Proposition 2.1.2 of \\cite{Ber1990}, a field $K_r$ containing $k$ which is flat over $k$ with respect to the completed tensor product and non-trivially valued. When we take the tensor product of the canonical morphism \\[\\mathcal{A}_{V}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{B}\\to \\mathcal{A}_{V}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B}\\] with $K_{r}$ over $k$ we get, using the flatness of $K_r$, that\n\\[(\\mathcal{A}_{V} \\widehat{\\otimes}_{k} K_r)\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A} \\widehat{\\otimes}_{k} K_r}(\\mathcal{B} \\widehat{\\otimes}_{k} K_r) \\to \\mathcal{A}_{V} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B} \\widehat{\\otimes}_{k} K_r\n\\] is an isomorphism in $\\sD^{-}(\\mathcal{A}\\widehat{\\otimes}_{k}K_r)$ and hence the original morphism $\\mathcal{A}_{V}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{A}_{V}\\to \\mathcal{A}_{V} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B}$ is an isomorphism in $\\sD^{-}(\\mathcal{A}).$\n\n\\ \\hfill $\\Box$\n\\begin{lem}Let $\\mathcal{A}_{W_1}$ and $\\mathcal{A}_{W_2}$ be affinoid localizations of an affinoid $k$-algebra $\\mathcal{A}$ corresponding to subdomains $W_1$ and $W_2$. Assume also that $W_1 \\cup W_2$ is an affinoid subdomain. Let $\\mathcal{B}$ be an affinoid $k$-algebra. Assume that the morphisms \n\\[\\mathcal{A}_{W_i}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{B} \\to \\mathcal{A}_{W_i}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B}\n\\]\nare isomorphisms for $i=1,2$. Assume also that the morphism\n\\[\\mathcal{A}_{W_1 \\cap W_2}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{B} \\to \\mathcal{A}_{W_1 \\cap W_2}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B}\n\\]\nis an isomorphism. Then the morphism \n\\[\\mathcal{A}_{W_1 \\cup W_2}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{B} \\to \\mathcal{A}_{W_1 \\cup W_2}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B}\n\\]\nis an isomorphism.\n\\end{lem}\n{\\bf Proof.}\nThis follows immediately from considering the strict short exact sequence \n\\[0 \\to \\mathcal{A}_{W_1 \\cup W_2} \\to \\mathcal{A}_{W_1} \\times \\mathcal{A}_{W_2} \\to \\mathcal{A}_{W_1 \\cap W_2} \\to 0\n.\\]\n\n\\ \\hfill $\\Box$\nFinally, we are able to show in the following theorem that affinoid subdomains of affinoids give examples homotopy monomorphisms of affine schemes in the abstract sense.\n\\begin{thm} \\label{thm:TensLocs}Let $\\mathcal{A}_{V}$ be an affinoid localization of an affinoid $k$-algebra $\\mathcal{A}$. Let $\\mathcal{B}$ be an affinoid $k$-algebra. Then the natural morphism \n\\[\\mathcal{A}_{V}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{B} \\to \\mathcal{A}_{V}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B} \\]\nis an isomorphism in $\\sD^{-}(\\mathcal{A})$.\nIn particular let $\\mathcal{A}_{W_1}$ and $\\mathcal{A}_{W_2}$ be affinoid localizations of an affinoid algebra $\\mathcal{A}$. Then $\\mathcal{A}_{W_1} \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}} \\mathcal{A}_{W_2} \\cong \\mathcal{A}_{W_1} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{W_2}.$ Therefore taking $W_1=W_2$, any affinoid localization is a homotopy epimorphism.\n\\end{thm}\n\n{\\bf Proof.}\nFor Weirstrass or Laurent localizations this follows immediately from using induction on Lemma \\ref{lem:easyLW} or the fact that it holds for the more general set of rational localizations which we now consider. When $\\mathcal{A}_{W_1}$ and $\\mathcal{A}_{W_2}$ are rational localizations this has been shown already in Lemma \\ref{lem:LCSrHoEpis}. Suppose now that $V=W_1$ is an rational domain and $W=W_2=V_1 \\cup \\cdots \\cup V_{N}$ is an affinoid domain written as a union of rational domains. For $N=1$ the claim is true. The induction step follows from considering the derived tensor product of $\\mathcal{A}_{V}$ over $\\mathcal{A}$ with the short exact sequence \n\\begin{equation}\\label{equation:ind}0\\to \\mathcal{A}_{W} \\to \\mathcal{A}_{V_1 \\cup \\cdots \\cup V_{N-1}} \\times \\mathcal{A}_{V_N} \\to \\mathcal{A}_{V_1 \\cup \\cdots \\cup V_{N-1}} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_N} \\to 0.\n\\end{equation}\nAssume by induction that the for some $N>2$ the derived tensor product of the rational localization $\\mathcal{A}_V$ and the affinoid algebra corresponding to the union of $N-1$ or fewer rational domains is equivalent to the ordinary tensor product. This implies that the derived tensor product of $\\mathcal{A}_{V}$ with $\\mathcal{A}_{V_1 \\cup \\cdots \\cup V_{N-1}} \\times \\mathcal{A}_{V_N}$ and with $\\mathcal{A}_{V_1 \\cup \\cdots \\cup V_{N-1}} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_N} =\\mathcal{A}_{(V_1 \\cap V_{N})\\cup \\dots \\cup (V_{N-1} \\cap V_{N})}$ are equivalent to the ordinary tensor products. Hence the same holds with $\\mathcal{A}_{W}$. Finally, one takes $W$ as $W_1$ in (\\ref{equation:ind}) and considers the derived tensor product with $\\mathcal{A}_{W_2}$ and uses a similar induction to do the general case.\n\n\n\\ \\hfill $\\Box$\n\n\n\n\\begin{rem}\\label{rem:recover}This is analogous to Lemma 2.1.4 (1) of \\cite{TVe5} where it is shown that homotopy Zariski open immersions in the category of affine schemes relative to the closed symmetric monoidal category of abelian groups give precisely the ordinary notion of a Zariski open imemrsions. In that reference first a special case (inverting a single element of the ring) was shown and the general case follows by their descent formalism.\n\\end{rem}\n\\begin{defn}\nFor any topological space $X$, a {\\it quasi-net} is a set $T$ of subsets of $X$ such that any point $x \\in T$ has a neighborhood of the form $\\cup_{i=1}^{n}V_{i}$ with $x \\in V_{i} \\in T$ for $1 \\leq i \\leq n$. A {\\it net} is a quasi-net $T$ such that such that for every $U,V \\in T$ the set $\\{W \\in T | W \\subset U \\cap V\\}$ is a quasi-net of subsets of $U \\cap V$. A $k$-analytic space is a locally Hausdorff topological space $X$, a net $\\tau_{0}$ on $X$, a functor $\\phi: \\tau_{0} \\to {\\text{\\bfseries\\sf{Afnd}}}_{k},$ and an invertible natural transformation ${\\text{\\bfseries\\sf{Top}}} \\Longrightarrow {\\text{\\bfseries\\sf{Top}}} \\circ \\phi$. \n\\end{defn}\n\\begin{defn} \nFor any $k$-affinoid algebra $\\mathcal{A}$, the topological space $|\\mathcal{M}(\\mathcal{A})|$ is defined to be the set of non-archimedean bounded semivaluations $| \\ |$ on $\\mathcal{A}$ equipped with the weakest topology such that for each $f \\in \\mathcal{A}$, the maps $|\\mathcal{M}(\\mathcal{A})| \\to \\mathbb{R}_{+}$ defined by sending $| \\ |$ to $|f|$ is continuous.\n\\end{defn}\n\n\\begin{defn}A {\\it k-affinoid space} is a locally ringed space of the form $\\mathcal{M}(\\mathcal{A})=(|\\mathcal{M}(\\mathcal{A})|,\\mathcal{O}_{\\mathcal{M}(\\mathcal{A})})$ where $\\mathcal{A}$ is a $k$-affinoid algebra and $\\mathcal{O}_{\\mathcal{M}(\\mathcal{A})}(U)$ is the limit over $\\mathcal{A}_{V}$, where $V \\subset U$ is a finite union of affinoid domains. The category of $k$-affinoid spaces defined to be a full subcategory of the category of locally ringed spaces of the given form. \n\\end{defn}\nThe category of $k$-affinoid spaces is equivalent to the category ${\\text{\\bfseries\\sf{Afnd}}}_{k}^{op}$. So we treat $\\mathcal{M}$ as a functor giving this equivalence from ${\\text{\\bfseries\\sf{Afnd}}}_{k}^{op}$ to the category of $k$-affinoid spaces.\n\\begin{defn}\nA {\\it k-analytic space} consists of a triple $(X,\\tau,\\mathcal{A})$ where $X$ is a locally Hausdorff topological space, $\\tau$ is a net on $X$, and for each $V \\in \\tau$, $\\mathcal{A}(V)$ is a k-affinoid algebra along with a homeomorphism $|\\mathcal{M}(\\mathcal{A}(V))| \\cong V$ (functorially assigned to the elements of $\\tau$) such that if $V, V' \\in \\tau$ and $V' \\subset V$ then $V'$ is an affinoid subdomain of $\\mathcal{M}(\\mathcal{A}(V))$ with coordinate ring $\\mathcal{A}(V')= \\mathcal{A}(V)_{V'}$. In the event that for every $U \\in \\tau_{2},$ $\\tau_{1}$ restricted to $g^{-1}(U)$ is an atlas of $g^{-1}(U),$ a morphism $(X_{1},\\tau_{1},\\mathcal{A}_{1}) \\to (X_{2},\\tau_{2},\\mathcal{A}_{2})$ consists of a continuous and $G$-continuous map $g:X_{1} \\to X_{2}$ along with bounded homomorphisms $g^{\\#}_{U,V}:\\mathcal{A}_{2}(U) \\to \\mathcal{A}_{1}(V)$ for every $U \\in \\tau_{2}, V \\in \\tau_{1}$ with $g(V) \\subset U$ such that for every $V, V' \\in \\tau_{1}$ with $V'\\subset V$ and $U, U' \\in \\tau_{2}$ with $U' \\subset U$ such that $g(V) \\subset U$ and $g(V') \\subset U'$ the diagram \n\\begin{equation}\n\\xymatrix{ \\mathcal{A}_{2}(U) \\ar[d] \\ar[r] & \\mathcal{A}_{1}(V) \\ar[d] \\\\ \\mathcal{A}_{2}(U') \\ar[r] & \\mathcal{A}_{1}(V') }\n\\end{equation}\ncommutes. We use the terms $k$-analytic space and Berkovich analytic spaces interchangeably. Let ${\\text{\\bfseries\\sf{An}}}_{k}$ denote the category of $k$-analytic spaces.\n\\end{defn}\n\n\n\\begin{defn}\\label{quasinet}A quasi-net on a topological space $X$ is a collection $T$ of subsets of $X$ such that for every $x \\in X$ there is a subset $T_x \\subset T$ with $|T_x| < \\infty$ such that $x \\in \\cap_{V \\in T_x}V$ and there is an open set $U \\subset X$ such that $x \\in U \\subset \\cup_{V \\in T_x} V.$\n\\end{defn}\n\\begin{lem}\\label{easyquasinet}Any finite set $T=\\{V_1,V_2, \\dots V_m \\}$ of closed subsets of a topological space $X$ which cover $X$ is a quasi-net.\n\\end{lem}\n{\\bf Proof.} Given $x \\in X,$ consider the subset $T_x \\subset T$ defined by those subsets in $T$ which contain $x$. Then $x \\in \\cap_{V \\in T_x}V$. Let $U=X-\\cup_{V \\in T-T_x}V,$ this satisfies the required property.\n\\ \\hfill $\\Box$\n\nConsider ${\\text{\\bfseries\\sf{Afnd}}}^{op}_{k}\/X$, the category of affinoid $k$-analytic spaces over $X$. This means that the objects are pairs $(\\mathcal{M}(\\mathcal{A}),f)$ where $\\mathcal{M}(\\mathcal{A})$ is an affinoid $k$-analytic space and $f:\\mathcal{M}(\\mathcal{A}) \\to X$ is a morphism of $k$-analytic spaces. The morphisms from $(\\mathcal{M}(\\mathcal{A}_1),f_1)$ to $(\\mathcal{M}(\\mathcal{A}_2),f_2)$ are morphisms to $X$ commuting with the $f_i$. \n\\begin{lem}\\label{EverythingColimit} Any $k$-analytic space $X$ is a colimit of the category ${\\text{\\bfseries\\sf{Afnd}}}^{op}_{k}\/X$ when considered as a subcategory of ${\\text{\\bfseries\\sf{An}}}_{k}$.\n\\end{lem}\n{\\bf Proof.}\nConsider the family $\\hat{\\tau}$ of all affinoid domains in $|X|$. It is a net and $|X|$ has a maximal $k$-affinoid atlas $\\hat{\\mathcal{A}}$. For each $V \\in \\hat{\\tau}$, $\\hat{\\mathcal{A}}$ assigns $\\mathcal{M}(\\mathcal{A}_{V})\\to X.$ In particular, $\\hat{\\mathcal{A}}$ assigns homeomorphisms $|\\mathcal{M}(\\mathcal{A}_{V})| \\cong V\\subset|X|$ and such that these homeomorphisms satisfy obvious compatibilities. For any other $k$-analytic space $X'$ we have by Exercise 3.2.2 of \\cite{Ber2009} an isomorphism\n\\begin{equation}\n\\Hom(X,X') \\rightarrow \\eq[\\prod_{V \\in \\hat{\\tau}} \\Hom(\\mathcal{M}(\\mathcal{A}_{V}), X')\\rightrightarrows \\prod_{(V,W) \\in \\hat{\\tau}^{2}} \\Hom(\\mathcal{M}(\\mathcal{A}_{V}\\widehat{\\otimes_{k}}\\mathcal{A}_{W}), X')\n].\\end{equation}\nTogether with the factorization of this isomorphism as\n\\footnotesize\n\\[\\Hom(X,X') \\to \\lim_{\\mathcal{M} \\in {\\text{\\bfseries\\sf{Afnd}}}^{op}_{k}\/X}\\Hom(\\mathcal{M},X') \\hookrightarrow \\eq[\\prod_{V \\in \\hat{\\tau}} \\Hom(\\mathcal{M}(\\mathcal{A}_{V}), X')\\rightrightarrows \\prod_{(V,W) \\in \\hat{\\tau}^{2}} \\Hom(\\mathcal{M}(\\mathcal{A}_{V}\\widehat{\\otimes_{k}}\\mathcal{A}_{W}),X')]\n\\]\n\\normalsize\nthis implies that the natural morphism \n\\[\\Hom(X,X') \\to \\lim_{\\mathcal{M} \\in {\\text{\\bfseries\\sf{Afnd}}}^{op}_{k}\/X}\\Hom(\\mathcal{M},X')\n\\]\nis an isomorphism. Therefore, $X=\\colim_{\\mathcal{M} \\in {\\text{\\bfseries\\sf{Afnd}}}^{op}_{k}\/X} \\mathcal{M}.$\n\\ \\hfill $\\Box$\n\\subsection{From Banach algebraic geometry to Berkovich geometry}\\label{BSSOCBA}\nConsider the category $\\text{\\bfseries\\sf{C}}={\\text{\\bfseries\\sf{Ban}}}_{k}$ for some valuation field $k$. We have shown in Section \\ref{BanachSpaces} that it is a closed symmetric monoidal quasi-abelian categories with $\\underline{\\Hom}=\\underline{\\Hom}_{k}$ and $\\overline{\\otimes} =\\widehat{\\otimes}_{k}$ with all finite limits and colimits and enough projectives so that we can do algebraic geometry relative to ${\\text{\\bfseries\\sf{Ban}}}_{k}$. In particular the categories of affine schemes over it has certain distinguished morphisms and topologies and we have notions of (Archimedean\/non-Archimedean) Banach schemes Banach (infinity) stacks and $n$-algebraic Banach stacks over an (Archimedean\/non-Archimedean) valuation field $k$. \nIn the Archimedean case we could compare this geometry to the geometry of complex varieties covered by Stein compact subsets. However, we focus here on the non-Archimedean case. In this section $k$ will be a non-Archimedean valuation field.\n\n\\begin{defn}\\label{def:NDiskrelA}\nLet $\\mathcal{A}$ be a $k$-affinoid algebra. Given $r_{1}, \\cdots, r_{n} \\in \\mathbb{R}$, we can define an $\\mathcal{A}$ algebra \n\\begin{equation}\\mathcal{A}\\{r_1^{-1}T_1, \\dots, r_n^{-1}T_n\\}\n\\end{equation}\nas the completion of $\\mathcal{A}[T_1, \\dots, T_n]$ with respect to the norm \n\\[\\|\\sum a_{I} T^{I}\\|_{r}=\\text{max}_{I}\\{\\|a_{I}\\|_{\\mathcal{A}}r^{I}\\}.\n\\] \n\\end{defn}\nLet $r$ be a real number greater than zero, denote by $\\mathcal{A}_{r}$ the $\\mathcal{A}$ module with norm $\\|a\\|= r\\|a\\|_{\\mathcal{A}}$.\n\\begin{lem}\\label{lem:DiskAlg}$\\mathcal{A}\\{r_1^{-1}T_1, \\dots, r_n^{-1}T_n\\}$ is the symmetric algebra (see subsection \\ref{Symmetric}) on $\\mathcal{V}=\\mathcal{A}_{r_1} \\oplus \\cdots \\oplus \\mathcal{A}_{r_n}$ in ${\\text{\\bfseries\\sf{Mod}}}^{\\leq 1}(\\mathcal{A})$. It can also be seen as the filtered colimit in ${\\text{\\bfseries\\sf{Mod}}}^{\\leq 1}(\\mathcal{A})$ of \n\\[\\sS^{0}(\\mathcal{V}) \\hookrightarrow \\sS^{0}(\\mathcal{V}) \\oplus \\sS^{1}(\\mathcal{V}) \\hookrightarrow \\sS^{0}(\\mathcal{V}) \\oplus \\sS^{1}(\\mathcal{V}) \\oplus \\sS^{2 }(\\mathcal{V})\\hookrightarrow \\cdots \n\\]\nwhere \n\\[\\sS^{m}(\\mathcal{V}) = \\{ \\sum_{|I|=m} a_{I} T^{I} | a_{I} \\in \\mathcal{A}\\}\\]\nequipped with the norm \n\\[\\|\\sum_{|I|=m} a_{I} T^{I} \\|= \\max_{|I|=m} \\|a_I\\| r^{I}.\n\\]\n\\end{lem}\n{\\bf Proof.} Left to the reader.\n\\ \\hfill $\\Box$\n\n\n\n\n\\begin{lem}\\label{lem:FinTypePres} \n\nA morphism $p: \\mathcal{A} \\to \\mathcal{B}$ in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}^{\\leq 1}_{k})$ induces a presentation \n\\begin{equation}\\label{eqn:NiceForm} \\mathcal{B} \\cong \\mathcal{A}\\{\\frac{T_1}{r_1}, \\dots, \\frac{T_g}{r_g}\\}\/(P_1, \\dots, P_r) \n\\end{equation}\nwhere $P_{i} \\in \\mathcal{A}\\{\\frac{T_1}{r_1}, \\dots, \\frac{T_g}{r_g}\\}$\nif and only if $p$ is of finite presentation in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}^{\\leq 1}_{k})$ as defined in Definition \\ref{defn:FinitePres}.\n\\end{lem}\n\n\\ \\hfill $\\Box$\n\nThe following two lemmas are technical results that will be used only in the proof of Theorem \\ref{thm:localForm}.\n\\begin{lem}\\label{SplitProdStr2}Let $\\mathcal{A}, \\mathcal{C}$ be $k$-affinoid algebras considered as objects in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}_{k})$ and let $f:\\mathcal{A} \\to \\mathcal{C}$ be morphism in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}_{k})$ which is a strict epimorphism when considered in ${\\text{\\bfseries\\sf{Ban}}}_{k}$ such that there is a $k$-affinoid algebra $\\mathcal{B}$ and the post-composition of $f$ with some homotopy epimorphism $g:\\mathcal{C} \\to \\mathcal{B}$ in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}_{k})$ is a homotopy epimorphism $h:\\mathcal{A}\\to \\mathcal{B}$ in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}_{k})$. Then there is a $k$-affinoid algebra $\\mathcal{A}'$ and a isomorphism $\\mathcal{A} \\cong \\mathcal{C} \\times \\mathcal{A}'$ such that the projection to $\\mathcal{C}$ corresponds to $f$ under this isomorphism.\\end{lem}\n{\\bf Proof.} \nThe morphisms \n\\[\\mathcal{A}\\to \\mathcal{C} \\to \\mathcal{B}\n\\]\ninduce morphisms on the derived categories \n\\[\\sD^{-}(\\mathcal{B}) \\to \\sD^{-}(\\mathcal{C}) \\to \\sD^{-}(\\mathcal{A}).\n\\]\nSince the composition is fully faithful and the first morphism is as well, the second morphism must be fully faithful and so by Lemma \\ref{lem:HomotopyMon} we have $\\mathcal{C} \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}} \\mathcal{C} \\cong \\mathcal{C}$. Let $I=\\ker (f)$. There is a strict, short exact sequence \n\\begin{equation}\\label{eqn:idealAC} 0 \\to I \\to \\mathcal{A} \\to \\mathcal{C} \\to 0.\n\\end{equation}\nIf we consider the derived completed tensor product of (\\ref{eqn:idealAC}) over $\\mathcal{A}$ with $\\mathcal{C}$ we find an exact triangle \n\\[I \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}} \\mathcal{C} \\to \\mathcal{C} \\to \\mathcal{C} \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}} \\mathcal{C}\n\\]\nand because the second morphism is an isomorphism, we see that \n$I \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}} \\mathcal{C}$ is isomorphic to $0$. If we now consider the derived completed tensor product over $\\mathcal{A}$ of (\\ref{eqn:idealAC}) with $I$ we an exact triangle \n\\[I \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}} I \\to I \\to I \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}} \\mathcal{C}\n\\]\nand so we get an isomorphism $I \\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}I \\to I$.\nSo we have $I=\\text{image}[I \\widehat{\\otimes}_{\\mathcal{A}}I \\to I]$ and in fact this implies that $I= I^2:= \\text{image}[I \\otimes_{\\mathcal{A}}I \\to I]$. Therefore, there exists an element $e \\in \\mathcal{A}$ such \nthat $e^{2}=e$ and $e\\mathcal{A} =I$. This gives the structure of a $k$-affinoid algebra \nto $I$, which we denote \nby $\\mathcal{A}'=\\mathcal{A}\/(1-e)\\mathcal{A}$. Now because $f$ is \na strict epimorphism, there is a strict short exact sequence \n\\[0 \\to I \\to \\mathcal{A} \\stackrel{f}\\to \\mathcal{C} \\to 0\n\\]\nwhich in fact is split by the morphism of algebras $e:\\mathcal{A} \\to \\mathcal{A}'.$ Therefore, \\[(e,f):\\mathcal{A} \\to \\mathcal{A}' \\times \\mathcal{C}\\] is \nan isomorphism.\n\\ \\hfill $\\Box$\n\\begin{lem}\\label{CoverOfAff}Let $\\mathcal{A}, \\mathcal{B}$ be $k$-affinoid algebras and let $f:\\mathcal{A} \\to \\mathcal{B}$ be a morphism in the category of $k$-affinoid algebras with the property that $|\\mathcal{M}(\\mathcal{A})|$ has a finite covering by affinoid domains $V_{j}$ corresponding to affinoid domain embeddings $\\mathcal{M}(\\mathcal{A}_{V_j})\\to \\mathcal{M}(\\mathcal{A}).$ Suppose also that morphisms $\\mathcal{M}(\\mathcal{A}_{V_j}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B}) \\to \\mathcal{M}(\\mathcal{A}_{V_j})$ are affinoid domain embeddings. Then the morphism $\\mathcal{M}(\\mathcal{B}) \\to \\mathcal{M}(\\mathcal{A})$ is an affinoid domain embedding.\n\\end{lem}\n{\\bf Proof.}\nLet us denote by $U$ the image of $|\\mathcal{M}(\\mathcal{B})|$ inside $|\\mathcal{M}(\\mathcal{A})|.$ We have $\\overline{U} \\cap V_{j} = \\overline{U \\cap V_j} $ for all $j$. Since $U \\cap V_j$ is closed in $V_j$ and $V_j$ is closed in $|\\mathcal{M}(\\mathcal{A})|$ we see that $U \\cap V_j$ is closed in $|\\mathcal{M}(\\mathcal{A})|$. Therefore $\\overline{U} \\cap V_{j} = U \\cap V_{j} $ for all $j$. Hence $\\overline{U}=U$ and so $U$ is closed inside $|\\mathcal{M}(\\mathcal{A})|.$\nLet $\\mathcal{A} \\to \\mathcal{C}$ be a bounded homomorphism of affinoid $k$-algebras such that the image of $|\\mathcal{M}(\\mathcal{C})|$ lies in $U$. We wish to show that the morphism $\\mathcal{M}(\\mathcal{C}) \\to \\mathcal{M}(\\mathcal{A})$ factors through a morphism $\\mathcal{M}(\\mathcal{C}) \\to \\mathcal{M}(\\mathcal{B})$. Notice that $\\mathcal{A}_{V_j} \\to \\mathcal{C} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} $ is a bounded homomorphism of affinoid $k$-algebras such that the image of $\\mathcal{M}(\\mathcal{C} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} )$ lies in $U\\cap V_j$. Therefore, the morphisms $\\mathcal{M}(\\mathcal{C} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} ) \\to \\mathcal{M}(\\mathcal{A}_{V_j} )$ factor in a unique way through morphisms $\\mathcal{M}(\\mathcal{C} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} ) \\to \\mathcal{M}(\\mathcal{A}_{V_j}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B}).$ When thought of as morphisms $\\mathcal{M}(\\mathcal{C} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} ) \\to \\mathcal{M}(\\mathcal{B})$ they agree when pulled back to $\\mathcal{M}(\\mathcal{C} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j \\cap V_k} ). $ The preimages of $V_j$ in $|\\mathcal{M}(\\mathcal{C})|$ are analytic domains by \\cite{Te} Exercise 3.2.2 (v). These preimages are the pullback of a quasi-net and therefore form a quasi-net by Lemma \\ref{easyquasinet} and therefore by Exercise 3.2.2 (v) of \\cite{Ber2009} we have a unique morphism $\\mathcal{M}(\\mathcal{C}) \\to \\mathcal{M}(\\mathcal{B})$ which restricts to the morphisms $\\mathcal{M}(\\mathcal{C} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} ) \\to \\mathcal{M}(\\mathcal{B}).$ \nIndeed, this follows from the commutative diagram of exact sequences \n\\footnotesize\n\\begin{equation}\n\\xymatrix{\n0 \\ar[r] & \\Hom(\\mathcal{M}(\\mathcal{C}),\\mathcal{M}(\\mathcal{B})) \\ar[d] \\ar[r] & \\prod_{j}\\Hom(\\mathcal{M}(\\mathcal{C}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} ),\\mathcal{M}(\\mathcal{B})) \\ar[r] \\ar[d] & \n\\prod_{j,k}\\Hom(\\mathcal{M}(\\mathcal{C}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j\\cap V_k} ),\\mathcal{M}(\\mathcal{B})) \\ar[d] \\\\ \n0 \\ar[r] &\\Hom(\\mathcal{M}(\\mathcal{C}),\\mathcal{M}(\\mathcal{A})) \\ar[r] & \\prod_{j}\\Hom(\\mathcal{M}(\\mathcal{C}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j} ),\\mathcal{M}(\\mathcal{A})) \\ar[r] & \\prod_{j,k}\\Hom(\\mathcal{M}(\\mathcal{C}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_j\\cap V_k} ),\\mathcal{M}(\\mathcal{A})). }\n\\end{equation}\n\\normalsize\nThis clearly provides the required factorisation. \n\\ \\hfill $\\Box$\n\nIf also, the $\\mathcal{M}(\\mathcal{A}_{V_j})$ are rational in $\\mathcal{M}(\\mathcal{A})$ and $\\mathcal{M}(\\mathcal{A}_{V_j}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B})$ is rational in $\\mathcal{M}(\\mathcal{A}_{V_j})$ notice that $\\mathcal{M}(\\mathcal{B})$ is a union of the rational domains $\\mathcal{M}(\\mathcal{A}_{V_j}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{B})$ in $\\mathcal{M}(\\mathcal{A}).$\n \n\\begin{lem}\\label{lem:NonExpFP} Let $\\mathcal{A}$ be a $k$-affinoid algebra and suppose there is a morphism $f: \\mathcal{A} \\to \\mathcal{B}$ of finite presentation in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}^{\\leq 1}_{k}).$ Then $\\mathcal{B}$ is a $k$-affinoid algebra.\n\\end{lem}\n{\\bf Proof.}\nBy combining a presentation for $\\mathcal{B}$ over $\\mathcal{A}$ and a presentation for $\\mathcal{A}$ over $k$ one can write $\\mathcal{B}$ as a finite colimit of objects of finite presentation in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}^{\\leq 1}_{k})$. Therefore, $\\mathcal{B}$ has finite presentation in ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}^{\\leq 1}_{k})$.\n\\ \\hfill $\\Box$\n\\begin{thm}\\label{thm:localForm}Let $\\mathcal{A}, \\mathcal{B}$ be $k$-affinoid algebras and let $f:\\mathcal{A} \\to \\mathcal{B}$ be a morphism in the category of $k$-affinoid algebras. Assume that $f$ is a homotopy epimorphism (see Definition \\ref{defn:homotopyEpi}) when considered in the category ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}_{k})$. Then the morphism $\\mathcal{M}(\\mathcal{B}) \\to \\mathcal{M}(\\mathcal{A})$ corresponding to $f$ is an affinoid domain embedding.\\end{thm}\n{\\bf Proof.} \nWe refer here to Temkin's proof \\cite{Te2} of the Gerritzen-Grauert Theorem for morphisms of affinoid algebras. This theorem, assuming only the epimorphism condition on $f$, produces a finite collection of morphisms of $k$-affinoid algebas $\\mathcal{A} \\to \\mathcal{A}_{V_i}$ corresponding to rational domain embeddings $\\mathcal{M}(\\mathcal{A}_{V_i}) \\to \\mathcal{M}(\\mathcal{A})$ covering $|\\mathcal{M}(A)|$ with the images $V_i.$ The theorem further ensures that the morphisms $\\mathcal{A}_{V_i} \\to \\mathcal{B} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_i}$ induced from $f$ admit factorzations \\[\\mathcal{A}_{V_i} \\twoheadrightarrow \\mathcal{C}_{i} \\hookrightarrow \\mathcal{B} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_i}.\\] These factorizations correspond to the composition of the morphism of $k$-affinoid algebras $\\mathcal{C}_{i} \\to (\\mathcal{C}_{i})_{W_i}=\\mathcal{B} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_i}$ corresponding to Weierstrass domain embeddings with the surjective morphisms of affinoid $k$-algebras $\\mathcal{A}_{V_i} \\to \\mathcal{C}_{i}$ corresponding to closed immersions. \nTherefore, by Lemma \\ref{lem:LCSrHoEpis} the morphism $\\mathcal{C}_{i} \\to \\mathcal{B} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_i}$ is a homotopy epimorphism in the category ${\\text{\\bfseries\\sf{Comm}}}({\\text{\\bfseries\\sf{Ban}}}_{k})$. Notice that \n\\[(\\mathcal{B}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{A}_{V_i})\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}_{V_i}}(\\mathcal{B}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{A}_{V_i}) \\cong \\mathcal{B}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{A}_{V_i}\n\\]\nbecause homotopy epimorphisms are closed under derived base change. However, applying Lemma \\ref{lem:LCSrHoEpis} we have \n\\[\\mathcal{B}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{A}_{V_i} \\cong \\mathcal{B}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_i}\n\\]\nand so we see that \n\\[(\\mathcal{B}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_i})\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}_{V_i}}(\\mathcal{B}\\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_i}) \\cong \\mathcal{B}\\widehat{\\otimes}^{\\mathbb{L}}_{\\mathcal{A}}\\mathcal{A}_{V_i}\n\\]\nand so by Lemma \\ref{lem:HomotopyMon} the morphisms $\\mathcal{A}_{V_i} \\to \\mathcal{B} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_i}$ are homotopy epimorphisms. Also, the morphisms of affinoid algebras corresponding to Weierstrass domain embeddings are injective. Therefore, Lemma \\ref{SplitProdStr2} can applied by choosing the $f$ from that lemma to be the morphism $\\mathcal{A}_{V_i} \\to \\mathcal{C}_{i}$ and $g$ from that lemma to be the morphism $\\mathcal{C}_{i} \\to \\mathcal{B} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_i}.$ The lemma then tells us that the morphism $\\mathcal{M}(\\mathcal{C}_i) \\to \\mathcal{M}(\\mathcal{A}_{V_i})$ is simply the inclusion of a connected component in a disjoint union of affinoids. Therefore, $\\mathcal{M}(\\mathcal{C}_i) \\to \\mathcal{M}(\\mathcal{A}_{V_i})$ is an affinoid domain embedding. Because the composition of affinoid domain embeddings is an affinoid domain embedding, we conclude that the morphisms $\\mathcal{M}(\\mathcal{B} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_i})\\to \\mathcal{M}(\\mathcal{A}_{V_i})$ are affinoid domain embeddings as well. By Lemma \\ref{CoverOfAff} we conclude that the original morphism gives an affinoid domain embedding $\\mathcal{M}(\\mathcal{B}) \\to \\mathcal{M}(\\mathcal{A})$.\n\n\\ \\hfill $\\Box$\n\n\n\nFrom now on we write ${\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A})$ in place of ${\\text{\\bfseries\\sf{Mod}}}_{{\\text{\\bfseries\\sf{Afnd}}}_{k}^{op}}^{RR}(\\mathcal{A})$.\n\\begin{lem}\\label{lem:ConsImpliesSur}\nLet $\\mathcal{A}$ be a $k$-affinoid algebra. Let $\\{f_{i}:\\mathcal{A} \\to \\mathcal{A}_{V_i}\\}_{i \\in I}$ be a family of affinoid localizations such that for some finite set $J \\subset I$ the corresponding family of functors \n\\[{\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A}) \\to {\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A}_{V_i})\n\\]\nfor $i \\in J$ is conservative. Then the morphism $\\coprod_{i\\in J}\\mathcal{M}(\\mathcal{A}_{V_i}) \\to \\mathcal{M}(\\mathcal{A})$ is surjective.\n\\end{lem}\n{\\bf Proof.}\nWe argue by contradiction. First assume that $k$ is non-trivially valued and $\\mathcal{A}$ is strictly affinoid. Suppose that the family of functors is conservative and some point $x\\in \\mathcal{M}(\\mathcal{A})$ is not in the image. By Proposition 2.1.15 of \\cite{Ber1990} the subset of points of $y\\in \\mathcal{M}(\\mathcal{A})$ such that $\\ker(|\\;|_{y})$ is a maximal ideal is a dense subset of $\\mathcal{M}(\\mathcal{A})$. Therefore, since the image is the closed set $\\cup_{i\\in J}V_{i}$ we may assume (by changing the point $x$) that $x\\in \\mathcal{M}(\\mathcal{A})$ is not in the image and $\\ker(|\\;|_{x})$ is a maximal ideal. Chose using Proposition 2.2.3 (iii) of \\cite{Ber1990} an affinoid subdomain $W$ of $\\mathcal{M}(\\mathcal{A})$ such that $x \\in W$ and $W \\cap V_{i}$ is empty for all $i \\in J$. Consider the \n morphism $0 \\to \\mathcal{A}_W$ of ${\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A})$. It is not an isomorphism but for each $i \\in J$, the pullback to each $\\spec(\\mathcal{A}_{V_i})$ is the isomorphism $0 \\to 0=\\mathcal{A}_{W} \\widehat{\\otimes}_{A} \\mathcal{A}_{V_i}$ of ${\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A}_{V_i})$. This gives a contradiction. For the general case, choose using Proposition 2.1.2 of \\cite{Ber1990}, a valuation field extension $k \\to K$ such that the valuation on $K$ is non-trivial and $\\mathcal{A}\\widehat{\\otimes}_{k}K$ is a strictly $K$-affinoid algebra. Notice that the conservativity assumption on the original family implies by Lemma \\ref{cor:BaseChangeConservHzAR} applied to the base change $\\spec(\\mathcal{A}\\widehat{\\otimes}_{k}K) \\to \\spec(\\mathcal{A})$ that the family of functors $\\{{\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A}\\widehat{\\otimes}_{k}K) \\to {\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A}_{V_i}\\widehat{\\otimes}_{k}K)\\}_{i \\in J}$ is also conservative. The morphism $\\coprod_{i\\in J}\\mathcal{M}(\\mathcal{A}_{V_i}\\widehat{\\otimes}_{k} K) \\to \\mathcal{M}(\\mathcal{A}\\widehat{\\otimes}_{k}K)$ cannot be surjective because in the commutative diagram, \n\\[\n\\xymatrix{ \\coprod_{i\\in J}\\mathcal{M}(\\mathcal{A}_{V_i}\\widehat{\\otimes}_{k}K) \\ar[d] \\ar[r] & \\coprod_{i\\in J}\\mathcal{M}(\\mathcal{A}_{V_i}) \\ar[d] \\\\ \\mathcal{M}(\\mathcal{A}\\widehat{\\otimes}_{k}K) \\ar[r] & \\mathcal{M}(\\mathcal{A}) }\\]\nthe horizonal arrows are surjective. Therefore, we have reduced to the previous case and so the proof is complete.\n\\ \\hfill $\\Box$\n\n\\begin{lem}\\label{lem:AlternatingTate}Consider a (surjective) cover of $X= \\mathcal{M}(\\mathcal{A})$ by a finite collection of affinoid domains $V_i= \\mathcal{M}(A_{V_{i}})$. Then the complex \n\\[\n0 \\to \\mathcal{A} \\to \\prod_{i_1} \\mathcal{A}_{V_{i_1}} \\to \\prod_{i_1 < i_2} \\mathcal{A}_{V_{i_1}} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_{i_2}} \\to \\cdots \\to \\mathcal{A}_{V_{1}} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_{2}}\\widehat{\\otimes}_{\\mathcal{A}} \\cdots \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_{n}} \\to 0. \n\\]\nis strictly exact.\n\\end{lem}\n{\\bf Proof.} By Proposition 1, section 8.1 of \\cite{BGR}, the inclusion (which is strict) of alternating cochains inside all cochains is a quasi-isomorphism. On the other hand, the complex of all cochains is strictly exact by Proposition 2.2.5 of \\cite{Ber1990}. The conclusion follows.\n\\ \\hfill $\\Box$\n\\begin{lem}\\label{lem:CoversRconservative}Consider a (surjective) cover of $X= \\mathcal{M}(\\mathcal{A})$ by a finite collection of affinoid domains $V_i= \\mathcal{M}(A_{V_{i}})$. Then the corresponding family of functors ${\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A}) \\to {\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A}_{V_i})$ is conservative.\n\\end{lem}\n{\\bf Proof.}\nLet $f: \\mathcal{M} \\to \\mathcal{N}$ in ${\\text{\\bfseries\\sf{Mod}}}^{RR}(\\mathcal{A})$ be any morphism such that $f_{i}: \\mathcal{M} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_i} \\to \\mathcal{N} \\widehat{\\otimes}_{\\mathcal{A}}\\mathcal{A}_{V_i} $ are isomorphisms for all $i$. \nThe alternating version of the \\v{C}ech-Amitsur complex (see Definition \\ref{defn:Amitsur}) corresponding to the morphism $\\mathcal{A} \\to \\prod_{i=1}^{n} \\mathcal{A}_{V_i}$ is a strictly exact bounded above complex by \\ref{lem:AlternatingTate} and so defines an element of $D^{-}(\\mathcal{A})$ (in fact the $0$ element!).\n\\[0 \\to \\mathcal{A} \\to \\prod_{i_1} \\mathcal{A}_{V_{i_1}} \\to \\prod_{i_1 < i_2} \\mathcal{A}_{V_{i_1}} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_{i_2}} \\to \\cdots \\to \\mathcal{A}_{V_{1}} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_{2}}\\widehat{\\otimes}_{\\mathcal{A}} \\cdots \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_{n}} \\to 0. \n\\]\nEach object in this complex is acyclic for the functor $\\mathcal{M} \\widehat{\\otimes}_{\\mathcal{A}}(-)$ since $\\mathcal{M}$ (being RR-quasi-coherent) is transversal to localizations of $\\mathcal{A}$. Therefore by Proposition \\ref{prop:ProjFProj}, if we apply the derived functor $\\mathcal{M} \\widehat{\\otimes}_{\\mathcal{A}}^{\\mathbb{L}}(-)$ we are left with a strictly exact complex \n\\[0 \\to \\mathcal{M} \\to \\prod_{i_1} \\mathcal{M} \\widehat{\\otimes}_{\\mathcal{A}} \\mathcal{A}_{V_{i_1}} \\to \\prod_{i_10$.\n\n\\subsection{Carnot-Carath\\'eodory distance} The Carnot-Carath\\'eodory distance on $\\mathbb{H}^n$ is defined by \n\\begin{equation} \\label{e:cc-dist}\n d(x,y) = \\inf \\{ length_{g_0} (\\gamma); \\; \\gamma \\text{ horizontal } C^1 \\text{-smooth curve joining } x \\text{ to } y\\},\n\\end{equation}\nwhere a $C^1$-smooth curve is said to be horizontal if, at every point, its tangent vector belongs to the horizontal subbundle of the tangent bundle and $g_0$ is the left invariant Riemannian metric which makes $(X_1,\\dots,X_n,Y_1,\\dots,Y_n,T)$ an orthonormal basis. For a general presentation of Carnot-Carath\\'eodory spaces, see e.g. \\cite{bel}, \\cite{mont}. \n\nThe topology induced by this distance is the original (Euclidean) topology on $\\mathbb{H}^n \\thickapprox (\\mathbb{R}^{2n+1},g_0)$ and $(\\mathbb{H}^n,d)$ is a complete metric space. The distance is left invariant and 1-homogeneous with respect to the dilations,\n\\begin{equation*}\nd(x \\cdot y, x\\cdot z) = d(y,z) \\quad \\text{and} \\quad d(\\delta_r(y), \\delta_r(z)) = r\\,d(y,z)\n\\end{equation*}\nfor all $x$, $y$, $z\\in\\mathbb{H}^n$ and all $r>0$. It follows in particular that $B(x,r) = x\\cdot \\delta_r (B(0,1))$ and hence \n\\begin{equation} \\label{e:measball}\n \\mathcal{L}^{2n+1}(B(x,r)) = c_n \\, r^{2n+2}\n\\end{equation}\nfor all $x\\in \\mathbb{H}^n$, all $r>0$ and where $c_n := \\mathcal{L}^{2n+1}(B(0,1))>0$. The measure $\\mathcal{L}^{2n+1}$ is in particular a doubling measure on $(\\mathbb{H}^n,d)$. For more details about doubling metric measure spaces, see e.g. \\cite{heinonen}.\n\nEndowed with its Carnot-Carath\\'eodory distance $\\mathbb{H}^n$ is a geodesic space, i.e., for all $x$, $y\\in\\mathbb{H}^n$, there exists a curve $\\sigma \\in C([a,b],\\mathbb{H}^n)$ such that $\\sigma(a)=x$, $\\sigma(b)=y$ and $d(x,y) = l(\\sigma)$ where \n\\begin{equation*}\n l(\\sigma) = \\sup_{N\\in \\mathbb{N}^*} \\sup_{a=t_0\\leq \\dots \\leq t_N = b} \\sum_{i=0}^{N-1} d(\\sigma(t_i),\\sigma(t_{i+1})).\n\\end{equation*}\nUp to a reparameterization one can always assume that length minimizing curves $ \\sigma$ are parameterized proportionally to arc-length, i.e., \n\\begin{equation*}\nd(\\sigma(s),\\sigma(s')) = v \\,(s'-s)\n\\end{equation*}\nfor all $s0$, $x \\cdot \\sigma_{\\chi,-2\\pi}$ if $t<0$, for all $\\chi \\in \\mathbb{C}^n$ such that $|\\chi| = \\sqrt{\\pi |t|}$.\n\\end{enumerate}\n\\end{thm}\n\nHere and in the following, $|\\chi| = (\\sum_{j=1}^n |\\chi_j|^2)^{1\/2}$ for $\\chi = (\\chi_1,\\dots,\\chi_n)\\in\\mathbb{C}^n$. In particular it follows from this description that $(\\mathbb{H}^n,d)$ is non-branching.\n\n\\begin{prop} [Non-branching property of $\\mathbb{H}^n$] \\label{nonbranching}\n The space $(\\mathbb{H}^n,d)$ is non-branching, i.e., any two minimal curves which coincide on a non trivial interval coincide on the whole intersection of their intervals of definition. \n\\end{prop}\n\nEquivalently for any quadruple of points $z$, $x$, $y$, $y'\\in \\mathbb{H}^n$, if $z$ is a midpoint of $x$ and $y$ as well as a midpoint of $x$ and $y'$, then $y=y'$. \n\nThe next lemma collects some differentiability properties of the distance function to a given point to be used later. For $y\\in \\mathbb{H}^n$, we set $L_y:= y\\cdot L$.\n\n\\begin{lem} \\label{prop-distcc} Let $y\\in \\mathbb{H}^n$ and set $d_y(x) := d(x,y)$. Then the function $d_y$ is of class $C^{\\infty}$ on $\\mathbb{H}^n \\backslash L_y$ (equipped with the usual differential structure when identifying $\\mathbb{H}^n$ with $\\mathbb{R}^{2n+1}$). Moreover one has \n\n\\renewcommand{\\theenumi}{\\roman{enumi}}\n\\begin{enumerate}\n\n\\item $|\\nabla_H d_y(x)| = 1$ for all $x \\in \\mathbb{H}^n \\backslash L_y$ where $$\\nabla_H d_y(x) := (X_1 d_y(x)+i Y_1 d_y(x), \\dots, X_n d_y(x)+i Y_n d_y(x)).$$ \n\n\\item If $\\nabla d_y(x) = \\nabla d_{y'}(x)$ and $d(x,y) = d(x,y')$ for some $x \\in \\mathbb{H}^n \\backslash ( L_y\\cup L_{y'})$, then $y=y'$. Here $\\nabla = (\\partial_{\\xi_1},\\cdots,\\partial_{\\xi_n},\\partial_{\\eta_1},\\cdots,\\partial_{\\eta_n}, \\partial_t)$ denotes the classical gradient when identifying $\\mathbb{H}^n$ with $\\mathbb{R}^{2n+1}$. \n \\end{enumerate}\n\\end{lem}\n\n\\begin{proof} Set $\\Phi(\\chi,\\varphi):=\\sigma_{\\chi,\\varphi}(1)$ where $\\sigma_{\\chi,\\varphi}$ is given in Theorem~\\ref{geod}. This map is a $C^\\infty$-diffeomorphism from $\\mathbb{C}^n\\setminus\\{0\\} \\times (-2\\pi,2\\pi)$ onto $\\mathbb{H}^n\\setminus L$ (see e.g. \\cite{monti}, \\cite{ar}, \\cite{juillet}). If $x = \\Phi(\\chi,\\varphi) \\in \\mathbb{H}^n\\setminus L$ with $(\\chi,\\varphi) \\in \\mathbb{C}^n\\setminus\\{0\\} \\times (-2\\pi,2\\pi)$, one has $d_0(x) = |\\chi|$ and \n\\begin{equation*}\n\\nabla_H d_0(x) = \\dfrac{\\chi}{|\\chi|} e^{-i\\varphi} \\quad \\text{and} \\quad \\partial_t d_0(x) = \\dfrac{\\varphi}{4|\\chi|},\n\\end{equation*}\nsee \\cite[Lemma 3.11]{ar}. Next, by left invariance, we have $d_y(x) = d_0(y^{-1}\\cdot x)$, $\\nabla_H d_y(x) = \\nabla_H d_0(y^{-1}\\cdot x)$ and $\\partial_t d_y(x) = \\partial_t d_0(y^{-1}\\cdot x)$ if $x\\in \\mathbb{H}^n \\backslash L_y$ and the lemma follows easily.\n\\end{proof}\n\n\\subsection{Interpolation between measures} \\label{sect-interpolation} The notion of interpolation constructed from a transport plan between any two measures will be one of the key notion to be used later. To define it in our geometrical context, we first fix a measurable selection of minimal curves, i.e., a Borel map $S:\\mathbb{H}^n\\times\\mathbb{H}^n \\rightarrow C([0,1], \\mathbb{H}^n)$ such that for all $x$, $y\\in\\mathbb{H}^n$, $S(x,y)$ is a minimal curve joining $x$ and $y$. The existence of such a measurable recipe to join any two points in $\\mathbb{H}^n$ by a minimal curve follows from general theorems about measurable selections, see e.g. \\cite[Chapter 7]{villani}. Next we set $e_t(\\sigma) := \\sigma(t)$ for all $\\sigma \\in C([0,1],\\mathbb{H}^n)$ and $t\\in [0,1]$. In particular $e_t(S(x,y))$ denotes the point lying at distance $t \\, d(x,y)$ from $x$ on the selected minimal curve $S(x,y)$ between $x$ and $y$. \n\n\\begin{defi} \\label{interpolation} \n Let $\\mu$, $\\nu\\in \\mathcal P(\\mathbb{H}^n)$ and let $\\gamma \\in \\Pi(\\mu,\\nu)$. The interpolations between $\\mu$ and $\\nu$ constructed from $\\gamma$ are defined as the family $((e_t \\circ S)_\\sharp \\gamma))_{t\\in [0,1]}$ of Borel probability measures on $\\mathbb{H}^n$. \n\\end{defi}\n\nNote that these interpolations depend a priori on the measurable selection $S$ of minimal curves. This is actually not a serious issue for our purposes. We will moreover always consider interpolations constructed from transport plans that are concentrated on the set $\\Omega$ on which $S(x,y)$ is nothing but the unique minimal curve between $x$ and $y$. Note also for further reference that $S\\lfloor_\\Omega$ is continuous.\n\n\\subsection{Intrinsic differentiability} Intrinsic differentiability properties of real-valued Lipschitz functions on $\\mathbb{H}^n$, namely a Rademacher's type theorem, will be useful when considering 1-Lipschitz Kantorovich potentials. This theorem is a particular case of a more general result due to P. Pansu. We say that a group homomorphism $g :\\mathbb{H}^n \\rightarrow \\mathbb{R}$ is homogeneous if $g(\\delta_r(x)) = r\\,g(x)$ for \nall $x\\in\\mathbb{H}^n$ and all $r >0$. \n\n\\begin{defi}\n We say that a map $f:\\mathbb{H}^n \\rightarrow \\mathbb{R}$ is Pansu-differentiable at $x \\in \\mathbb{H}^n$ if there exists \nan homogeneous group homomorphism $g :\\mathbb{H}^n \\rightarrow \\mathbb{R}$ such that\n\\begin{equation*}\n\\lim_{y \\rightarrow x} \\frac{f(y) - f(x) - g(x^{-1} \\cdot y)}{d(y,x)} = 0.\n\\end{equation*}\nThe map $g$ is then unique and will be denoted by $D_H f(x)$.\n\\end{defi}\n\nIf $f:\\mathbb{H}^n \\rightarrow \\mathbb{R}$ is Pansu-differentiable at $x \\in \\mathbb{H}^n$ then the maps $s \\mapsto f(x \\cdot \\delta_s[e_j, 0])$, resp. $s \\mapsto f(x\\cdot \\delta_s[e_{n+j}, 0])$, are differentiable at $s=0$ and if we denote the corresponding derivatives by $X_jf(x)$, resp. $Y_jf(x)$, then \n\\begin{equation*}\nD_H f(x)(\\xi,\\eta,t) = \\sum_{j=1}^n \\xi_j X_jf(x) + \\eta_j Y_jf(x).\n\\end{equation*}\nHere $e_j = (\\delta_{1}^{j},\\dots,\\delta_{n}^{j})\\in \\mathbb{C}^n$ and $e_{n+j} = (i\\delta_{1}^{j},\\dots,i\\delta_{n}^{j})\\in \\mathbb{C}^n$. Using similar notations as in the classical smooth case, we then set $\\nabla_H f(x) := (X_1 f(x)+i Y_1 f(x), \\dots, X_n f(x)+i Y_n f(x))$.\n\n\\begin{thm}[Pansu-differentiability theorem] \\label{PansuRademacher} \\cite{pansu}\nLet $f:(\\mathbb{H}^n,d) \\rightarrow \\mathbb{R}$ be a $C$-Lipschitz function. Then, for $\\mathcal{L}^{2n+1}$-a.e. $x\\in\\mathbb{H}^n$, the function $f$ is Pansu-differentiable at $x$ and $|\\nabla_H f(x)|\\leq C$.\n\\end{thm}\n\nThe next lemma will be used to prove that any optimal transport plan is concentrated on the set $\\Omega$.\n\n\\begin{lem} \\label{uniquegeod}\n Let $u\\in \\text{Lip}_1(d)$, $x\\in\\mathbb{H}^n$ be such that $u$ is Pansu-differentiable at $x$ with $|\\nabla_H u(x)| \\leq 1$ and let $y\\in\\mathbb{H}^n$ be such that $u(x) - u(y) = d(x,y)$. Then there exists a unique minimal curve between $x$ and $y$.\n\\end{lem}\n\n\\begin{proof}\n Let $\\sigma:[0,1]\\rightarrow\\mathbb{H}^n$ be a minimal curve between $x$ and $y$. Then $\\sigma$ is a horizontal $C^1$-smooth curve and if $\\sigma(t) = (\\sigma_1(t),\\dots,\\sigma_{2n+1}(t)) \\in \\mathbb{H}^n \\thickapprox \\mathbb{R}^{2n+1}$, one has for all $t\\in [0,1]$,\n\\begin{equation*}\n \\dot\\sigma(t) = \\sum_{j=1}^n \\dot\\sigma_j(t)\\, X_j(\\sigma(t)) + \\dot\\sigma_{n+j}(t)\\, Y_j(\\sigma(t))\n\\end{equation*}\nand $|\\dot\\sigma_H(t)| = d(x,y)$ where $\\dot\\sigma_H(t) := (\\dot\\sigma_1(t) +i\\,\\dot\\sigma_{n+1}(t),\\dots,\\dot\\sigma_n(t) +i\\,\\dot\\sigma_{2n}(t))\\in\\mathbb{C}^n$. On the other hand, one has \n\\begin{equation*}\n u(x) - u(\\sigma(t)) = d(x,\\sigma(t)) = t\\, d(x,y)\n\\end{equation*}\nfor all $t\\in [0,1]$. Differentiating this equality with respect to $t$, we get\n\\begin{equation*}\n \\sum _{j=1}^n \\dot\\sigma_j(0)\\, X_ju(x) + \\dot\\sigma_{n+j}(0) \\, Y_ju(x)= \\dfrac{d}{dt}\\,u(\\sigma(t))_{|_{t=0}} = - d(x,y).\n\\end{equation*}\nAll together, it follows that \n\\begin{equation*}\nd(x,y) = |\\sum _{j=1}^n \\dot\\sigma_j(0)\\, X_ju(x) + \\dot\\sigma_{n+j}(0) \\, Y_ju(x)| \\leq |\\nabla_H u(x)|\\, |\\dot\\sigma_H(0)| \\leq d(x,y).\n\\end{equation*}\nIn particular, there is equality in all the previous inequalities which implies in turn that $\\dot\\sigma_H(0) = - \\,d(x,y) \\,\\nabla_H u (x)$. On the other hand one knows from Theorem~\\ref{geod} that $\\sigma = x\\cdot \\sigma_{\\chi,\\varphi}$ for some $\\chi \\in \\mathbb{C}^n\\setminus\\{0\\}$ and $\\varphi\\in\\, [-2\\pi,2\\pi]$. In particular one has $\\dot\\sigma_H(0) = \\chi$. It follows that $\\chi = - \\,d(x,y) \\,\\nabla_H u (x)$ is uniquely determined hence there is a unique minimal curve joining $x$ and $y$ according once again to the description given in Theorem~\\ref{geod}.\n\\end{proof}\n\n\n\\section{Properties of $\\Pi_1(\\mu,\\nu)$ and $\\Pi_2(\\mu,\\nu)$} \\label{sect:optplanning}\n\nLet $\\mu$, $\\nu \\in \\mathcal P_c(\\mathbb{H}^n)$ be fixed. We denote by $\\Pi_1(\\mu,\\nu)$ the set of optimal transport plans solution to Kantorovich transport problem~\\eqref{e:MK} between $\\mu$ and $\\nu$ with cost $c(x,y) = d(x,y)$. \n\nWe first prove some geometric properties of optimal transport plans. These properties follow from the behavior of minimal curves in $(\\mathbb{H}^n,d)$. In the next lemma, we prove that any optimal transport plan is concentrated on the set $\\Omega$ (see \\eqref{e:omega}) of pair of points that are connected by a unique minimal curve.\n\n\\begin{lem} \\label{pi1.1}\n Let $\\gamma\\in \\Pi_1(\\mu,\\nu)$ and assume that $\\mu\\ll\\mathcal{L}^{2n+1}$. Then for $\\gamma$-a.e. $(x,y)$, there exists a unique minimal curve between $x$ and $y$.\n\\end{lem}\n\n\\begin{proof}\nLet $u\\in \\text{Lip}_1(d)$ be a Kantorovich potential associated to Kantorovich transport problem~\\eqref{e:MK} between $\\mu$ and $\\nu$ with cost $c(x,y) = d(x,y)$ (see Section~\\ref{opttrans} and Theorem~\\ref{1lip_potential} there). Since $u\\in \\text{Lip}_1(d)$, we know from Theorem \\ref{PansuRademacher} that for $\\mathcal{L}^{2n+1}$-a.e., and hence $\\mu$-a.e., $x\\in\\mathbb{H}^n$, $u$ is Pansu-differentiable at $x$ with $|\\nabla_H u(x)|\\leq 1$. Then the conclusion follows from Lemma~\\ref{uniquegeod} since $u(x) - u(y) = d(x,y)$ for $\\gamma$-a.e. $(x,y)$ (see Theorem~\\ref{1lip_potential}). \n\\end{proof}\n\nThe next lemma says that minimal curves used by an optimal transport plan cannot bifurcate. It follows essentially from the non-branching property of $(\\mathbb{H}^n,d)$.\n\n\\begin{lem} \\label{pi1.2}\n Let $\\gamma\\in \\Pi_1(\\mu,\\nu)$. Then $\\gamma$ is concentrated on a set $\\Gamma$ such that the following holds. For all $(x,y)\\in \\Gamma$ and $(x',y')\\in \\Gamma$ such that $x\\not=y$ and $x\\not=x'$, if $x'$ lies on a minimal curve between $x$ and $y$ then all points $x$, $x'$, $y$ and $y'$ lie on the same minimal curve. More precisely, there exists a minimal curve $\\sigma:[a,b]\\rightarrow\\mathbb{H}^n$ such that $x=\\sigma(a)$, $y=\\sigma(t)$ for some $t\\in \\,(a,b]$, $x'=\\sigma(s)$ for some $s\\in\\, (a,t]$ and $y'=\\sigma(t')$ for some $t'\\in [s,b]$.\n\\end{lem}\n\n\\begin{proof} Let $(x,y)\\in \\mathbb{H}^n \\times \\mathbb{H}^n$ and $(x',y')\\in \\mathbb{H}^n \\times \\mathbb{H}^n$ such that $x\\not=y$ and $x'\\not=x$. Assume that $x'\\in \\sigma((0,d(x,y)])$ where $\\sigma:[0,d(x,y)] \\rightarrow \\mathbb{H}^n$ is a unit-speed minimal curve between $x$ and $y$. Let $\\sigma'$ be a unit-speed minimal curve between $x'$ and $y'$ parameterized on $[d(x,x'),d(x,x')+d(x',y')]$. Assume moreover that \n\\begin{equation*} \n d(x,y) + d(x',y') \\leq d(x,y')+d(x',y).\n\\end{equation*}\nRecall that this holds true for $\\gamma$-a.e. $(x,y)$ and $(x',y')$ by Theorem \\ref{ccycl}. Then the curve $\\tilde \\sigma :[0, d(x,x')+d(x',y')]\\rightarrow\\mathbb{H}^n$ which coincides with $\\sigma$ on $[0,d(x,x')]$ and $\\sigma'$ on $[d(x,x'),d(x,x')+d(x',y')]$ is a length minimizing curve between $x$ and $y'$. Indeed, otherwise we would have \n\\begin{equation*}\n d(x,y') < l(\\tilde\\sigma) = l(\\sigma_{|[0,d(x,x')]}) + l(\\sigma'_{|[d(x,x'),d(x,x')+d(x',y')]}) = d(x,x') + d(x',y').\n\\end{equation*}\nSince $x'$ lies on a minimal curve between $x$ and $y$, we have $d(x,x') + d(x',y) = d(x,y)$ and we get \n\\begin{equation*}\n d(x,y') + d(x',y) < d(x,y) + d(x',y')\n\\end{equation*}\nwhich gives a contradiction. It follows that $\\sigma$ and $\\tilde\\sigma$ are unit-speed minimal curves that coincide on the non trivial interval $[0,d(x,x')]$. Since $\\mathbb{H}^n$ is non-branching (see Proposition~\\ref{nonbranching}), this implies that $\\sigma$ and $\\tilde \\sigma$ are sub-arcs of the same minimal curve, namely $\\sigma$ if $d(x,y') \\leq d(x,y)$ and $\\tilde\\sigma$ otherwise, on which all points $x$, $x'$, $y$ and $y'$ lie. And the conclusion follows.\n\\end{proof}\n\nWe denote by $\\Pi_2(\\mu,\\nu)$ the set of transport plans solution to the secondary variational problem:\n\\begin{equation*}\n\\min_{\\gamma \\in \\Pi_1(\\mu,\\nu)} \\int_{\\mathbb{H}^n\\times\\mathbb{H}^n} d(x,y)^2\\,d\\gamma(x,y).\n\\end{equation*}\n\nOptimal transport plans selected through the variational approximations to be introduced in Section~\\ref{varapprox} will be solution to this secondary variational problem. The next lemma gives a one-dimensional monotonicity condition along minimal curves used by optimal transport plans in $\\Pi_2(\\mu,\\nu)$. This follows essentially from a constrained version of $d^2$-cyclical monotonicity.\n\n\\begin{lem} \\label{pi2}\nLet $\\gamma\\in \\Pi_2(\\mu,\\nu)$. Then $\\gamma$ is concentrated on a set $\\Gamma$ such that the following holds. For all $(x,y)\\in \\Gamma$ and $(x',y')\\in \\Gamma$ such that $x\\not=y$ and $x\\not=x'$, if $x'$ lies on a minimal curve between $x$ and $y$ then all points $x$, $x'$, $y$ and $y'$ lie on the same minimal curve ordered in that way.\n\\end{lem}\n\nIn other words, there exists a minimal curve $\\sigma:[a,b]\\rightarrow\\mathbb{H}^n$ such that $\\sigma(a) = x$, $\\sigma(t)=y$ for some $t\\in \\,(a,b]$, $\\sigma(s) = x'$ for some $s\\in\\, (a,t]$ and $\\sigma(t') = y'$ for some $t'\\in [t,b]$.\n\n\\begin{proof} First, as a classical fact, one can rephrase the secondary variational problem as a classical Kantorovich transport problem \\eqref{e:MK} between $\\mu$ and $\\nu$ with cost $c(x,y) = \\beta(x,y)$ with \n\\begin{equation*}\n\\beta(x,y) = \\begin{cases}\n d(x,y)^2 \\quad \\text{if } u(x)-u(y) = d(x,y),\\\\\n +\\infty \\qquad \\phantom{\\text{if}} \\text{otherwise},\n \\end{cases}\n\\end{equation*}\nwhere $u\\in \\text{Lip}_1(d)$ is a Kantorovich potential associated to Kantorovich transport problem~\\eqref{e:MK} between $\\mu$ and $\\nu$ with cost $c(x,y) = d(x,y)$ (see Section~\\ref{opttrans} and Theorem~\\ref{1lip_potential} there). Since $\\beta$ is lower semicontinuous and $\\int_{\\mathbb{H}^n\\times \\mathbb{H}^n} \\beta(x,y) \\,d\\gamma(x,y)<+\\infty$ for all $\\gamma \\in \\Pi_2(\\mu,\\nu)$, it follows from Theorem \\ref{ccycl} that any $\\gamma \\in \\Pi_2(\\mu,\\nu)$ is concentrated on a $\\beta$-cyclically monotone set. So, taking into account the fact that $\\Pi_2(\\mu,\\nu)\\subset\\Pi_1(\\mu,\\nu)$, we know that $\\gamma \\in \\Pi_2(\\mu,\\nu)$ is concentrated on a set $\\Gamma$ such that \n\\begin{equation*}\n u(x) - u(y) = d(x,y)\n\\end{equation*}\nfor all $(x,y)\\in \\Gamma$,\n\\begin{equation*}\n \\beta(x,y) +\\beta(x',y') \\leq \\beta(x,y')+\\beta(x',y)\n\\end{equation*}\nfor all $(x,y)\\in \\Gamma$ and $(x',y')\\in \\Gamma$ and the conclusion of Lemma \\ref{pi1.2} holds.\n\nThen let $(x,y)\\in \\Gamma$ and $(x',y')\\in \\Gamma$ be as in the statement. By Lemma \\ref{pi1.2}, the conclusion will follow if we show that $d(x',y) \\leq d(x',y')$. First we check that $\\beta(x',y) = d(x',y)^2$ and $\\beta(x,y') = d(x,y')^2$. We have \n\\begin{equation*}\nu(x) \\leq u(x') + d(x,x') \\leq u(y) + d(x',y) + d(x,x') = d(x,y) + u(y) = u(x)\n\\end{equation*}\nhence all these inequalities are equalities. In particular, we get that $u(x') = u(y) + d(x',y)$ hence $\\beta(x',y) = d(x',y)^2$. We also get that \n\\begin{equation*}\n u(x) = d(x,x') + u(x') = d(x,x') + u(y') + d(x',y') = u(y') + d(x,y')\n\\end{equation*}\nhence $\\beta(x,y') = d(x,y')^2$. If $d(x',y')< d(x',y)$, we get\n\\begin{equation*}\n\\begin{split}\n \\beta(x,y') + &\\beta(x',y) - \\beta(x',y') - \\beta(x,y)\\\\\n&= d(x,y')^2 + d(x',y)^2 - d(x',y')^2 - d(x,y)^2 \\\\\n&= (d(x,x')+d(x',y'))^2 + d(x',y)^2 - d(x',y')^2 - (d(x,x')+d(x',y))^2\\\\\n&= 2 \\, d(x,x') (d(x',y') - d(x',y))<0\n\\end{split}\n\\end{equation*}\nwhich gives a contradiction.\n\\end{proof}\n\n\n\\section{Variational approximations} \\label{varapprox}\n\nWe introduce variational approximations in the spirit of \\cite{ap} (see also \\cite{cfm}, \\cite{akp}) by rephrasing in our geometrical context the variational approximations considered recently in \\cite{sant}. This approximation procedure will be used to select optimal transport plans that will be eventually proved to be induced by transport maps.\n\nLet $\\mu$, $\\nu \\in \\mathcal P_c(\\mathbb{H}^n)$ be fixed. Let $K$ be a compact subset of $\\mathbb{H}^n$ such that $\\operatorname{spt}{\\mu}\\cup\\operatorname{spt}{\\nu}\\subset K$ and set\n\\begin{equation*}\n \\Pi := \\{\\gamma \\in \\mathcal P(\\mathbb{H}^n\\times\\mathbb{H}^n);\\, (\\pi_1)_\\sharp\\gamma=\\mu,\\, \\operatorname{spt}{(\\pi_2)_\\sharp\\gamma} \\subset K\\}.\n\\end{equation*}\nFor $\\varepsilon>0$ fixed and $\\gamma \\in \\Pi$, we set\n\\begin{multline*}\n C_\\varepsilon(\\gamma) := \n\\frac{1}{\\varepsilon}\\, W_1((\\pi_2)_\\sharp\\gamma,\\nu) + \\int_{\\mathbb{H}^n\\times\\mathbb{H}^n} d(x,y)\\,d\\gamma(x,y) \\\\ + \\varepsilon \\int_{\\mathbb{H}^n\\times\\mathbb{H}^n} d(x,y)^2\\,d\\gamma(x,y) + \\varepsilon^{6n+8} \\operatorname{card}{(\\operatorname{spt}{(\\pi_2)_\\sharp\\gamma})}\n\\end{multline*}\nand consider the family of minimization problems:\n\\begin{equation} \\label{e:varapprox}\n \\min\\{ C_\\varepsilon(\\gamma);\\, \\gamma \\in \\Pi\\}. \\tag{$P_\\varepsilon$}\n\\end{equation}\nHere $W_1$ denotes the 1-Wasserstein distance defined for any two probability measures $\\mu_1$, $\\mu_2 \\in \\mathcal P(\\mathbb{H}^n)$ by \n\\begin{equation*}\n W_1(\\mu_1,\\mu_2) := \\min_{\\gamma \\in \\Pi(\\mu_1,\\mu_2)} \\int_{\\mathbb{H}^n\\times\\mathbb{H}^n} d(x,y)\\,d\\gamma(x,y).\n\\end{equation*}\nFirst we note that \\eqref{e:varapprox} always admits solutions.\n\n\\begin{thm}\n For any $\\varepsilon>0$, the problem \\eqref{e:varapprox} admits at least one solution and $\\min\\{ C_\\varepsilon(\\gamma);\\, \\gamma \\in \\Pi\\}<+\\infty$.\n\\end{thm}\n\n\\begin{proof}\nFirst note that since $K$ is compact, $C_\\varepsilon(\\gamma)<+\\infty$ for any $\\gamma\\in \\Pi$ such that $(\\pi_2)_\\sharp\\gamma$ is finitely atomic. Next the existence of solutions to \\eqref{e:varapprox} follows from the weak compactness of $\\Pi$, the lower semicontinuity of the three first terms to be minimized and the Kuratowski convergence of the supports of weakly converging probability measures (see \\cite[Chapter~5]{ags}).\n\\end{proof}\n\nNext, weak limits of solutions to \\eqref{e:varapprox} are optimal transport plans that are solutions to the secondary variational problem introduced in Section~\\ref{sect:optplanning} to which we refer for the definition of $\\Pi_2(\\mu,\\nu)$. Modulo minor modifications due to our geometrical context, this can be proved with the same arguments as those given in \\cite{sant}. \n\n\\begin{lem} \\label{optpi2}\n Let $\\varepsilon_k$ be a sequence converging to 0 and $\\gamma_{\\varepsilon_k}$ a sequence of solutions to $(P_{\\varepsilon_k})$ which is weakly converging to some $\\gamma\\in \\mathcal P(\\mathbb{H}^n\\times\\mathbb{H}^n)$. Then $\\gamma \\in \\Pi_2(\\mu,\\nu)$.\n\\end{lem}\n\n\\begin{proof} First we note that for any $m\\geq 1$, one can find a finite set $F_m\\subset K$ such that $\\operatorname{card} F_m \\leq C\\, m^{2n+2}$ for some constant $C>0$ which depends only on $n$ and $\\operatorname{diam} K$ and a Borel map $p_m:K \\rightarrow F_m$ such that\n\\begin{equation*}\n d(p_m(x),x) < 1\/m \n\\end{equation*}\nfor all $x\\in K$. Indeed choose $x_1 \\in K$. For $i\\geq 2$, choose by induction $x_i \\in K\\setminus \\cup_{j0$ and set \n\\begin{equation*}\n c_\\varepsilon (x,y) := d(x,y) + \\varepsilon\\, d(x,y)^2.\n\\end{equation*}\nWe first recall the following classical fact.\n\n\\begin{lem} \\label{restrictions}\nLet $\\gamma_\\varepsilon$ be a solution to ($P_\\varepsilon$). Then for any Borel set $U\\subset\\mathbb{H}^n\\times\\mathbb{H}^n$, $(\\pi_2)_\\sharp(\\gamma_\\varepsilon\\lfloor U)$ is finitely atomic and $\\gamma_\\varepsilon\\lfloor U$ is a solution to Kantorovich transport problem \\eqref{e:MK} between $(\\pi_1)_\\sharp (\\gamma_\\varepsilon\\lfloor U)$ and $(\\pi_2)_\\sharp (\\gamma_\\varepsilon\\lfloor U)$ with cost $c_\\varepsilon$.\n\\end{lem}\n\n\\begin{proof}\n The fact that $(\\pi_2)_\\sharp(\\gamma_\\varepsilon\\lfloor U)$ is finitely atomic obviously follows from the fact that $C_\\varepsilon(\\gamma_\\varepsilon) = \\min\\{ C_\\varepsilon(\\gamma);\\, \\gamma \\in \\Pi\\}<+\\infty$. Next it is also immediate that $\\gamma_\\varepsilon$ is a solution to Kantorovich transport problem \\eqref{e:MK} between $\\mu$ and $(\\pi_2)_\\sharp (\\gamma_\\varepsilon)$ with cost $c_\\varepsilon$. Then as a classical fact, the claim follows from the linearity of the functional to be minimized with respect to the transport plan. If $\\gamma \\in \\Pi((\\pi_1)_\\sharp (\\gamma_\\varepsilon\\lfloor U),(\\pi_2)_\\sharp(\\gamma_\\varepsilon\\lfloor U))$, one indeed simply compare $C_\\varepsilon(\\gamma_\\varepsilon)$ with $C_\\varepsilon(\\hat\\gamma)$ where $ \\hat\\gamma = \\gamma_\\varepsilon\\lfloor (\\mathbb{H}^n\\times\\mathbb{H}^n)\\setminus U + \\gamma \\in \\Pi(\\mu,(\\pi_2)_\\sharp(\\gamma_\\varepsilon))$ to get the conclusion.\n\\end{proof}\n\nNext in this section we consider interpolations between two measures $\\overline\\mu$, $\\overline \\nu\\in \\mathcal P_c(\\mathbb{H}^n)$ that are constructed from a transport plan solution to Kantorovich transport problem \\eqref{e:MK} between these two measures with cost $c_\\varepsilon$. We prove absolute continuity and, more importantly, $L^\\infty$-estimates on the density with respect to $\\mathcal{L}^{2n+1}$ of these interpolations whenever $\\overline\\mu \\ll \\mathcal{L}^{2n+1}$ and $\\overline \\nu$ is finitely atomic, see Proposition~\\ref{e:Linftydensityestimates}. We divide the arguments into several steps. First we prove that any solution to this Kantorovich transport problem is induced by a transport.\n\n\\begin{thm} \\label{existenceceps}\n Let $\\overline\\mu$, $\\overline \\nu\\in \\mathcal P_c(\\mathbb{H}^n)$ be fixed. Assume that $\\overline\\mu \\ll \\mathcal{L}^{2n+1}$ and that $\\overline \\nu$ is finitely atomic. Then any solution to Kantorovich transport problem \\eqref{e:MK} between $\\overline\\mu$ and $\\overline \\nu$ with cost $c_\\varepsilon$ is induced by a transport. In particular there exists a unique optimal transport map solution to the transport problem \\eqref{e:M} between $\\overline\\mu$ and $\\overline \\nu$ with cost $c_\\varepsilon$.\n\\end{thm}\n\n\\begin{proof} Let $\\psi$ be a Kantorovich potential for Kantorovich transport problem \\eqref{e:MK} between $\\overline\\mu$ and $\\overline \\nu$ with cost $c_\\varepsilon$ given by Theorem~\\ref{duality}. Let $\\{y_i\\}_{i=1}^k$ denote the atoms of $\\overline \\nu$. We prove that for $\\mathcal{L}^{2n+1}$-a.e. $x\\in\\mathbb{H}^n$, there is at most one point $y_i$ for some $i\\in\\{1,\\dots,k\\}$ such that\n\\begin{equation*} \n \\psi(x) + \\psi^c(y_i) = c_\\varepsilon (x,y_i).\n\\end{equation*}\nSince $\\overline\\mu \\ll \\mathcal{L}^{2n+1}$, it will follow that any transport plan solution to Kantorovich transport problem \\eqref{e:MK} between $\\overline\\mu$ and $\\overline \\nu$ with cost $c_\\varepsilon$ is concentrated on a $\\overline\\mu$-measurable graph and hence induced by a transport. This implies in turn existence and uniqueness of the optimal transport map solution to the transport problem \\eqref{e:M} between $\\overline\\mu$ and $\\overline \\nu$ with cost $c_\\varepsilon$ (see Theorem~\\ref{plan-transport}). \n\nFor $i\\not= j$, set $h_{ij}(x) := c_\\varepsilon (x,y_i) - c_\\varepsilon (x,y_j) + \\psi^c(y_j) - \\psi^c(y_i)$. It follows from Lemma~\\ref{prop-distcc} that $h_{ij}$ is of class $C^\\infty$ on the open set $\\mathbb{H}^n\\setminus(L_{y_i} \\cup L_{y_j})$ with $\\nabla h_{ij} \\not= 0$. Indeed assume on the contrary that $\\nabla h_{ij}(x) =0$ for some $x\\in \\mathbb{H}^n\\setminus(L_{y_i} \\cup L_{y_j})$. Then, differentiating along the horizontal vector fields $X_j$ and $Y_j$, we would have\n\\begin{equation*} \n \\nabla_H d_{y_i} (x) \\,(1+2\\varepsilon\\, d_{y_i}(x)) = \\nabla_H d_{y_j} (x) \\,(1+2\\varepsilon \\,d_{y_j}(x)).\n\\end{equation*}\nSince $|\\nabla_H d_{y_i} (x) | = |\\nabla_H d_{y_j} (x) |$ (see Lemma~\\ref{prop-distcc}(i)), this would imply that $d_{y_i}(x) = d_{y_j}(x)$ and in turn that $\\nabla_H d_{y_i} (x) = \\nabla_H d_{y_j} (x)$. Since we also have by assumption $\\partial_t d_{y_i} (x) = \\partial_t d_{y_j} (x)$, Lemma~\\ref{prop-distcc}(ii) would give $y_i=y_j$. It follows that the set $\\{x\\in\\mathbb{H}^n\\setminus(L_{y_i} \\cup L_{y_j});\\, h_{ij}(x) =0\\}$ is a $C^\\infty$-smooth submanifold of dimension $2n$ in $\\mathbb{R}^{2n+1}$ and hence has Lebesgue measure 0. Since $\\mathcal{L}^{2n+1}(L_{y_i}) =0$, it follows that \n\\begin{equation*} \n \\mathcal{L}^{2n+1} (\\bigcup_{i\\not=j} \\{x\\in\\mathbb{H}^n;\\,h_{ij}(x) =0\\}) = 0\n\\end{equation*}\nand the claim follows.\n\\end{proof}\n\nIf $T:\\mathbb{H}^n \\rightarrow \\mathbb{H}^n$, we set $T_t = e_t \\circ S \\circ (I \\otimes T)$, i.e., $T_t(x)$ is the point lying at distance $t \\, d(x,T(x))$ from $x$ on the selected minimal curve $S(x,T(x))$ between $x$ and $T(x)$ (see Subsection~\\ref{sect-interpolation} for the definition of $S$ and $e_t$).\n\n\\begin{prop} \\label{injectivity} \\cite[Chapter 7]{villani}\n Let $\\overline\\mu$, $\\overline \\nu\\in \\mathcal P_c(\\mathbb{H}^n)$ be fixed such that $\\overline\\mu \\ll \\mathcal{L}^{2n+1}$ and $\\overline \\nu$ is finitely atomic. Let $T^\\varepsilon$ be the optimal transport map solution to the transport problem \\eqref{e:M} between $\\overline\\mu$ and $\\overline \\nu$ with cost $c_\\varepsilon$. Then there exists a $\\overline\\mu$ - measurable set $A$ such that $\\overline\\mu(A) = 1$ and such that for each $t\\in\\, [0,1)$, $T^\\varepsilon_t\\lfloor_A$ is injective.\n\\end{prop}\n\nThe cost $c_\\varepsilon$ can be recovered as coming from a so-called coercive Lagrangian action. Since $(\\mathbb{H}^n,d)$ is non-branching, the proposition essentially follows from~\\cite[Chapter 7, Theorem 7.30]{villani}. However one does not need the full strength of the theory developed in~\\cite[Chapter 7]{villani} to get the conclusion of Proposition~\\eqref{injectivity} and we sketch below the arguments for the reader's convenience. \n\n\\begin{proof} Let $0\\leq s0$, there exist $y'\\in\\mathbb{H}^n$ and $r'>0$ such that\n\n\\renewcommand{\\theenumi}{\\roman{enumi}}\n\\begin{enumerate}\n \\item $y \\in B(y',r') \\subset\\subset B(y,r)$,\n\n \\item $x\\in \\operatorname{Leb}{\\rho}$ and $\\rho(x) <+\\infty$,\n\n \\item $x\\in \\operatorname{Leb}{\\rho'}$ and $\\rho'(x)>0$,\n\\end{enumerate}\nwhere $\\rho$ denotes the density of $(\\pi_1)_\\sharp\\gamma$ and $\\rho'$ the density of $(\\pi_1)_\\sharp\\gamma\\lfloor(\\mathbb{H}^n\\times B(y',r'))$ with respect to $\\mathcal{L}^{2n+1}$.\n\\end{lem}\n\n\\begin{proof} Let $(y_m)_{m\\geq 1}$ be a dense sequence in $\\mathbb{H}^n$.\nFor each $m,k \\in \\mathbb{N}^*$, set $\\gamma_{m,k}:=\\gamma \\lfloor (\\mathbb{H}^n \\times B(y_m, r_k))$ where $r_k:=1\/k$. Let $\\rho_{m,k}$ denote the density of $(\\pi_1)_\\sharp\\gamma_{m,k}$ with respect to $\\mathcal{L}^{2n+1}$. Set $A_{m,k} := \\mathbb{H}^n \\setminus (\\operatorname{Leb} \\rho \\cap \\operatorname{Leb} \\rho_{m,k} \\cap \\{\\rho < +\\infty\\})$. We have $\\mathcal{L}^{2n+1}(A_{m,k})=0$. Since $(\\pi_1)_\\sharp\\gamma \\ll \\mathcal{L}^{2n+1}$, it follows that $\\gamma(A_{m,k}\\times B(y_m, r_k)) \\leq (\\pi_1)_\\sharp\\gamma (A_{m,k}) =0$. Next\n \\begin{equation*}\n \\gamma(\\{\\rho_{m,k}=0\\} \\times B(y_m,r_k)) = (\\pi_1)_\\sharp\\gamma_{m,k}(\\{\\rho_{m,k}=0\\}) = 0.\n \\end{equation*}\nIt follows that $\\gamma(D_{m,k})=0$ for all $m,k \\in \\mathbb{N}^*$ where\n\\begin{equation*}\n D_{m,k}:=\\left[\\mathbb{H}^n \\setminus (\\operatorname{Leb} \\rho \\cap \\operatorname{Leb} \\rho_{m,k} \\cap \\{\\rho < +\\infty\\} \\cap \\{\\rho_{m,k}>0\\}) \\right] \\times B(y_m,r_k)\n\\end{equation*}\nhence $\\gamma(\\cup_{m,k} D_{m,k})=0$ and $\\gamma$ is concentrated on $\\mathbb{H}^n \\setminus \\cup_{m,k} D_{m,k}$. Then the conclusion follows noting that for each $(x,y) \\in \\mathbb{H}^n\\times\\mathbb{H}^n$ and $r>0$, one can find $m,k\\in \\mathbb{N}^*$ such that $y \\in B(y_m,r_k)\\subset \\subset B(y,r)$.\n\\end{proof}\n\nWe say that $x\\in E$ is a Lebesgue point of a Borel set $E$ if $x\\in \\operatorname{Leb} {\\rm 1\\mskip-4mu l}_E$, i.e., if $x\\in E$ and \n\\begin{equation*}\n \\lim_{r\\rightarrow 0} \\dfrac{\\mathcal{L}^{2n+1}(E\\cap B(x,r))}{\\mathcal{L}^{2n+1}(B(x,r))} = 1,\n\\end{equation*}\nand we denote by $\\operatorname{Leb} E:= \\operatorname{Leb} {\\rm 1\\mskip-4mu l}_E$ the set of all Lebesgue points of $E$. Note that $\\mathcal{L}^{2n+1}(E\\setminus \\operatorname{Leb} E) = 0$.\n\nThe next lemma together with Lemma~\\ref{mainlemma} and Lemma~\\ref{pi2} is one of the key ingredients of the proof of Theorem~\\ref{mainthmbis} and eventually of the existence of a solution to Monge's transport problem. It can be recovered as a consequence of Lemma~\\ref{dens1}. However, for sake of clarity, we state and prove it independently.\n\n\\begin{lem} \\label{dens2}\n Let $\\gamma\\in \\mathcal P(\\mathbb{H}^n\\times\\mathbb{H}^n)$ be such that $(\\pi_1)_\\sharp\\gamma \\ll \\mathcal{L}^{2n+1}$. Assume that $\\gamma$ is concentrated on a $\\sigma$-compact set $\\Gamma$. For $y\\in\\mathbb{H}^n$ and $r>0$, set \n\\begin{equation*}\n \\Gamma^{-1} (B(y,r)) = \\pi_1(\\Gamma \\cap (\\mathbb{H}^n \\times B(y,r))).\n\\end{equation*}\nThen $\\Gamma^{-1} (B(y,r))$ is a Borel set and $\\gamma$ is concentrated on a set $\\Gamma'\\subset\\Gamma$ such that for all $(x,y)\\in\\Gamma'$ and all $r>0$, $x\\in \\operatorname{Leb}{\\Gamma^{-1} (B(y,r)})$.\n\\end{lem}\n\n\\begin{proof}\n Since $\\Gamma$ is $\\sigma$-compact, $\\Gamma^{-1} (B(y,r))$ is also $\\sigma$-compact hence a Borel set. Set $A := \\{(x,y) \\in \\Gamma;\\,x\\notin \\operatorname{Leb}\\Gamma^{-1}(B(y,r)) \\text{ for some } r>0 \\}$\nand let us show that $\\gamma(A)=0$. For each $k\\in\\mathbb{N}^*$, consider a countable covering of \n$\\mathbb{H}^n$ by balls $(B(y_i^k,r_k))_{\\,i \\geq 1}$ of radius $r_k:=1\/(2k)$. If $(x,y) \\in \\Gamma$ and $x \\notin \\operatorname{Leb}\\Gamma^{-1}(B(y,r))$ then for any $k \\geq 1\/r$ and $y_i^k$ such that\n$d(y_i^k,y)0$, we have\n\\begin{equation*}\n \\liminf_{\\delta\\downarrow 0} \\, \\dfrac{\\mathcal{L}^{2n+1}(T(\\Gamma \\cap [B(x,\\frac{\\delta}{2}) \\times B(y,r)]) \\cap B(x,\\delta))}{\\mathcal{L}^{2n+1}(B(x,\\delta))}\\, > 0.\n\\end{equation*}\n\\end{lem}\n\nThe proof below follows the line of the proof of the similar property in~\\cite{champion-dePascale}. In our context it requires however some technical refinement. \n\n \\begin{proof} We consider the set $\\Gamma$ obtained by Lemma~\\ref{dens1}, $(x,y) \\in \\Gamma$ with $x \\neq y$ and $r>0$. Then let $y'\\in\\mathbb{H}^n$ and $r'>0$ be given by Lemma~\\ref{dens1} so that Lemma~\\ref{dens1}(i), (ii) and (iii) hold. Using the same notations as in this lemma, we set \\begin{equation*}\n G:=\\{z \\in \\mathbb{H}^n;\\, \\ \\frac{1}{2}\\rho'(x) \\leq \\rho' (z) \\,\\, \\mbox{and} \\,\\,\n\\rho(z) \\leq 2\\rho(x) \\}.\n\\end{equation*}\nThen $G$ is a Borel set. We have $0< \\rho'(x) \\leq \\rho(x)$ (remember the convention about densities of absolutely continuous measure, see~\\eqref{e:density}). Since $x\\in \\operatorname{Leb} \\rho \\cap \\operatorname{Leb} \\rho'$, see Lemma~\\ref{dens1}(ii) and (iii), it follows that $x\\in\\operatorname{Leb} G$.\n\nFix $\\delta>0$ such that $\\delta < d(x,y)+r$ and \n\\begin{equation} \\label{e:reg-delta}\n \\frac{1}{2}\\mathcal{L}^{2n+1}(B(x,s)) \\leq \\mathcal{L}^{2n+1}(G \\cap B(x,s))\n\\end{equation}\nfor all $s \\in (0,\\delta)$ and fix $t>0$ such that $4t(d(x,y)+r)< \\delta$.\n\nWe set $G_\\delta:=G\\cap B(x,\\frac{\\delta}{2})$, $A_\\delta:=G_\\delta \\times B(y',r')$ and \n$\\gamma_{\\delta}:= \\gamma \\lfloor A_\\delta$.\nWe shall prove that\n\\begin{equation} \\label{e:main1}\n \\frac{\\rho'(x)}{4} \\, \\mathcal{L}^{2n+1}(B(x,\\frac{\\delta}{2})) \\leq (e_t \\circ S)_\\sharp \\gamma_{\\delta} (B(x,\\delta)) \n\\end{equation} \nand\n\\begin{equation} \\label{e:main2}\n (e_t \\circ S)_\\sharp \\gamma_{\\delta} (B(x,\\delta)) \\leq 2^{2n+4} \\rho(x) \\mathcal{L}^{2n+1} (T(\\Gamma \\cap [ B(x,\\frac{\\delta}{2}) \\times B(y,r)]) \\cap B(x,\\delta)).\n\\end{equation}\nThen \\eqref{e:main1} and \\eqref{e:main2} will yield \n\\begin{equation*}\n2^{-(2n+6)} \\frac{\\rho' (x)}{\\rho(x)}\\, \\mathcal{L}^{2n+1}(B(x,\\frac{\\delta}{2})) \\leq \\mathcal{L}^{2n+1}(T(\\Gamma \\cap [ B(x,\\frac{\\delta}{2}) \\times B(y,r)]) \\cap B(x,\\delta))\n\\end{equation*}\nfor any $\\delta>0$ small enough which completes the proof.\n\nTo prove \\eqref{e:main1}, we note that $(\\pi_1)_\\sharp \\gamma_{\\delta} \\ll \\mathcal{L}^{2n+1}$ with density bounded below by $\\frac{1}{2} \\rho'(x) $ $\\mathcal{L}^{2n+1}$-a.e. on $G_\\delta$. Together with \\eqref{e:reg-delta}, it follows that\n\\begin{equation*}\n \\frac{\\rho'(x)}{4} \\, \\mathcal{L}^{2n+1}(B(x,\\frac{\\delta}{2})) \\leq (\\pi_1)_\\sharp \\gamma_{\\delta} (B(x,\\frac{\\delta}{2})).\n\\end{equation*}\nNext, by choice of $\\delta$ and $t$, we have $(e_t \\circ S)(z,w) \\in B(x,\\delta)$ for all $z \\in B(x,\\frac{\\delta}{2})$ and $w \\in B(y,r)$, hence\n\\begin{equation*}\n B(x,\\frac{\\delta}{2}) \\times B(y',r') \\subset B(x,\\frac{\\delta}{2}) \\times B(y,r) \\subset (e_t \\circ S)^{-1} (B(x,\\delta))\n\\end{equation*}\nand it follows that \n\\begin{equation*}\n (\\pi_1)_\\sharp \\gamma_{\\delta} (B(x,\\frac{\\delta}{2})) = \\gamma_{\\delta} (B(x,\\frac{\\delta}{2}) \\times B(y',r')) \\leq (e_t \\circ S)_\\sharp \\gamma_{\\delta} (B(x,\\delta))\n\\end{equation*}\nand this completes the proof of \\eqref{e:main1}.\n\nWe prove now \\eqref{e:main2}. By hypothesis, $\\gamma$ is a weak limit of solutions $\\gamma_k$ to $(P_{\\varepsilon_k})$ for some sequence $\\varepsilon_k$ converging to 0. For each fixed $k\\in\\mathbb{N}$, we apply Lemma~\\ref{restrictions} with $U = G_\\delta \\times \\mathbb{H}^n$ and Proposition~\\ref{e:Linftydensityestimates} with $\\overline \\mu = (\\pi_1)_\\sharp (\\gamma_k \\lfloor U)$ and $\\overline \\nu = (\\pi_2)_\\sharp (\\gamma_k \\lfloor U)$. Taking into account the fact that $(\\pi_1)_\\sharp (\\gamma_k \\lfloor U) = \\mu \\lfloor G_\\delta$, we get that $(e_t \\circ S)_\\sharp (\\gamma_k \\lfloor G_\\delta \\times \\mathbb{H}^n) \\ll \\mathcal{L}^{2n+1}$ with density in $L^\\infty$ and whose $L^\\infty$-norm is bounded by \n\\begin{equation} \\label{e:rhok}\n \\frac{1}{(1-t)^{2n+3}} \\, \\|\\rho \\lfloor_{G_\\delta} \\|_{L^\\infty} \\leq 2^{2n+4} \\rho(x).\n\\end{equation}\nNext we check that $(e_t \\circ S)_\\sharp (\\gamma_k \\lfloor G_\\delta \\times \\mathbb{H}^n)$ converges weakly to $(e_t \\circ S)_\\sharp (\\gamma \\lfloor G_\\delta \\times \\mathbb{H}^n)$. First it follows from Lemma \\ref{weakres} (to be proved below) that $\\gamma_k\\lfloor G_\\delta \\times \\mathbb{H}^n$ converges weakly to $\\gamma \\lfloor G_\\delta \\times \\mathbb{H}^n$. Then, noting that $\\gamma$ and each $\\gamma_k$ are concentrated on $\\Omega$ and that $e_t \\circ S$ is continuous on $\\Omega$, the claim follows from Lemma \\ref{conv} (to be proved below) applied with $\\overline\\gamma = \\gamma \\lfloor G_\\delta \\times \\mathbb{H}^n$, $\\overline\\gamma_k = \\gamma_k \\lfloor G_\\delta \\times \\mathbb{H}^n$, $B=\\Omega$ and $f = \\varphi \\circ e_t \\circ S$ where $\\varphi \\in C_b(\\mathbb{H}^n)$. The fact that $\\gamma$ is concentrated on $\\Omega$ follows from Lemma \\ref{pi1.1}. To check that $\\gamma_k$ is concentrated on $\\Omega$, denote by $\\{y_i^k\\}_i$ the finite set of the atoms of $(\\pi_2)_\\sharp \\gamma_k$. We have that $\\gamma_k$ is concentrated on $\\mathbb{H}^n\\times \\{y_i^k\\}_i$. On the other hand $\\gamma_k(L_{ y_i^k} \\times \\{y_i^k\\}) \\leq \\gamma_k(L_{ y_i^k} \\times \\mathbb{H}^n) = \\mu(L_{ y_i^k}) = 0$ since $\\mu \\ll \\mathcal{L}^{2n+1}$. It follows that $\\gamma_k$ is concentrated on $\\cup_i [(\\mathbb{H}^n \\setminus L_{ y_i^k}) \\times \\{y_i^k\\}] \\subset \\Omega$. Then, taking into account \\eqref{e:rhok}, we get\n\\begin{equation*}\n |\\int_{\\mathbb{H}^n} \\varphi \\,\\, d(e_t \\circ S)_\\sharp (\\gamma \\lfloor G_\\delta \\times \\mathbb{H}^n)| \\leq 2^{2n+4} \\rho(x) \\, \\|\\varphi\\|_{L^1}\n\\end{equation*}\nfor every $\\varphi \\in C_b(\\mathbb{H}^n)$. It follows that $(e_t \\circ S)_\\sharp (\\gamma \\lfloor G_\\delta \\times \\mathbb{H}^n)$ is in $(L^1)'$ with density in $L^\\infty$ and whose $L^\\infty$-norm is bounded by $2^{2n+4} \\rho(x)$. Since $(e_t \\circ S)_\\sharp \\gamma_\\delta \\leq (e_t \\circ S)_\\sharp (\\gamma \\lfloor G_\\delta \\times \\mathbb{H}^n)$, the same holds true for $(e_t \\circ S)_\\sharp \\gamma_\\delta$. Finally we note that $\\gamma_\\delta$ being concentrated on $\\Gamma \\cap [ B(x,\\frac{\\delta}{2}) \\times B(y',r')] \\subset \\Gamma \\cap [ B(x,\\frac{\\delta}{2}) \\times B(y,r)]$, the measure $(e_t \\circ S)_\\sharp \\gamma_\\delta$ is concentrated on $T(\\Gamma \\cap [ B(x,\\frac{\\delta}{2}) \\times B(y',r')]) \\subset T(\\Gamma \\cap [ B(x,\\frac{\\delta}{2}) \\times B(y,r)])$. All together we get\n\\begin{equation*}\n\\begin{split}\n (e_t \\circ S)_\\sharp \\gamma_{\\delta} (B(x,\\delta)) &= (e_t \\circ S)_\\sharp \\gamma_{\\delta} (T(\\Gamma \\cap [ B(x,\\frac{\\delta}{2})\n\\times B(y,r)]) \\cap B(x,\\delta) )\\\\\n&\\leq 2^{2n+4} \\rho(x) \\mathcal{L}^{2n+1} (T(\\Gamma \\cap [ B(x,\\frac{\\delta}{2}) \\times B(y,r)]) \\cap B(x,\\delta))\n\\end{split}\n\\end{equation*}\nwhich proves \\eqref{e:main2}.\n\\end{proof}\n\n\\begin{lem} \\label{weakres} \nLet $X$ be a separable and locally compact Hausdorff metric space in which every open set is $\\sigma$-compact. Let $(\\gamma_k)_k$ be a sequence in $\\mathcal P (X\\times X)$\nwhich converges weakly to some $\\gamma \\in \\mathcal P (X\\times X)$ and\nsuch that $(\\pi_1)_\\sharp \\gamma_k = (\\pi_1)_\\sharp \\gamma$ for every $k\\in \\mathbb{N}$. Then for any Borel set $G \\subset X$, the sequence \n$(\\gamma_k \\lfloor G\\times X)_k$ converges weakly to $\\gamma \\lfloor G\\times X$.\n\\end{lem}\n\n\\begin{proof} We have to prove that for any $\\varphi \\in C_b(X)$,\n\\begin{equation*}\n\\lim_{k\\rightarrow +\\infty} \\int_{X\\times X} {\\rm 1\\mskip-4mu l}_G(x) \\varphi (x,y) \\,d \\gamma_k(x,y) = \\int_{X \\times X} {\\rm 1\\mskip-4mu l}_G(x) \\varphi(x,y) \\,d \\gamma(x,y).\n\\end{equation*}\n\nIt follows from Lusin's Theorem that for any \n$\\varepsilon>0$ there exists a closed set $F_\\varepsilon$ such that ${{\\rm 1\\mskip-4mu l}_G} \\lfloor_{F_\\varepsilon}$ is continuous and $(\\pi_1)_\\sharp \\gamma(X \\setminus\nF_\\varepsilon) < \\varepsilon$. As a consequence, for every $\\varepsilon >0$, the restriction of $(x,y) \\mapsto {\\rm 1\\mskip-4mu l}_G(x) \\varphi (x,y)$ to $F_\\varepsilon \\times X$ is continuous and \n$$\\limsup_{k\\rightarrow +\\infty} \\gamma_k ((X\\setminus F_\\varepsilon) \\times X)\n= (\\pi_1)_\\sharp \\gamma (X \\setminus F_\\varepsilon) < \\varepsilon.$$\nThen since $(x,y)\\mapsto |{\\rm 1\\mskip-4mu l}_G(x) \\varphi (x,y)|$ is\nbounded and hence uniformly integrable with respect to\n$(\\gamma_k)_k$, the claim follows from \\cite[Proposition 5.1.10]{ags}.\n\\end{proof}\n\n\\begin{lem} \\label{conv}\nLet $X$ be a separable metric space and $(\\overline\\gamma_k)_k$ be a sequence in $\\mathcal P(X)$ which converges weakly \nto some $\\overline\\gamma \\in\\mathcal P(X)$. Let $f:X\\to \\mathbb{R}$ be a measurable and bounded function which is continuous \nin $B$ for some Borel set $B\\subset X$ such that $\\overline\\gamma_k(X\\setminus B)=0$ for every $k\\in \\mathbb{N}$ and $\\overline\\gamma(X\\setminus B)=0$, then\n\\begin{equation*}\n \\lim_{k \\to \\infty} \\int_X f d\\gamma_k= \\int_X f d\\gamma.\n\\end{equation*}\n\\end{lem}\n\n\\begin{proof} Let $\\overline{f}$ and $\\tilde{f}$ be respectively the lower and\n upper semicontinuous envelope of $f$. We have $\\overline{f}=f=\\tilde{f}$ on $B$ and hence $\\gamma$-a.e. and $\\gamma_k$-a.e. for every $k\\in\\mathbb{N}$. It follows that\n\\begin{multline*}\n \\int_X f \\, d\\gamma = \\int_X \\overline f \\, d\\gamma \\leq \\liminf_{k \\to \\infty} \\int_X \\overline{f} \\,d\\gamma_k \n= \\liminf_{k \\to \\infty} \\int_X f \\,d\\gamma_k \\\\\n\\leq \\limsup_{k \\to \\infty} \\int_X f d\\gamma_k \n= \\limsup_{k \\to \\infty} \\int_X \\tilde f d\\gamma_k \\leq \\int_X \\tilde f d\\gamma = \\int_X f \\, d\\gamma\n\\end{multline*}\nwhich proves the claim.\n\\end{proof}\n\n\\section{Solution to Monge's problem} \\label{conclusion}\n\nWe prove that optimal transport plans in $\\Pi_1(\\mu,\\nu)$ that are obtained as weak limit of solutions of the variational approximations introduced in Section~\\ref{varapprox} are induced by a transport, hence giving a solution to Monge's transport problem as stated in Theorem~\\ref{mainthm}. Note that due to the fact that $\\Pi$ is relatively compact in $\\mathcal P(\\mathbb{H}^n\\times\\mathbb{H}^n)$, such optimal transport plans do exist.\n\n\\begin{thm} \\label{mainthmbis}\n Let $\\varepsilon_k$ be a sequence converging to 0 and $\\gamma_{\\varepsilon_k}$ a sequence of solutions to $(P_{\\varepsilon_k})$ which is weakly converging to some $\\gamma\\in \\mathcal P(\\mathbb{H}^n\\times\\mathbb{H}^n)$. Then $\\gamma$ is concentrated on a $\\mu$-measurable graph and hence induced by a transport.\n\\end{thm}\n\n\\begin{proof}\nFirst we know from Lemma~\\ref{optpi2} that $\\gamma\\in \\Pi_2(\\mu,\\nu)$. From the previous sections and using inner regularity of Borel probability measures, one can then find $\\sigma$-compact sets $\\Gamma$ and $\\Gamma'$ such that $\\Gamma'\\subset \\Gamma \\subset \\Omega$ and the conclusions of Lemma~\\ref{pi2}, Lemma~\\ref{dens2} and Lemma~\\ref{mainlemma} hold. We prove here that for any $x\\in \\pi_1(\\Gamma')$ there is a unique $y\\in \\mathbb{H}^n$ such that $(x,y)\\in \\Gamma'$.\n\nBy contradiction, assume that one can find $x_0 \\in \\pi_1(\\Gamma')$ and $(x_0,y_0)\\in \\Gamma$, $(x_0,y_1)\\in \\Gamma$ with $y_0\\not=y_1$. Without loss of generality one can assume that $d(x_0,y_0) \\leq d(x_0,y_1)$ and $x_0\\not=y_1$. Then, by Lemma \\ref{dens2} and Lemma \\ref{mainlemma}, for all $r>0$ and for all $\\delta>0$ small enough, one can find $x'\\in B(x_0,\\delta) \\cap \\Gamma^{-1} (B(y_0,r)) \\cap T(\\Gamma \\cap [B(x_0,\\frac{\\delta}{2}) \\times B(y_1,r)])$. It follows that one can find $y' \\in B(y_0,r)$ such that $(x',y') \\in \\Gamma$ and $(x,y) \\in \\Gamma \\cap (B(x_0,\\frac{\\delta}{2})\\times B(y_1,r))$ such that $x\\not=y$, $x'\\not=x$ and $x'$ lie on the minimal curve between $x$ and $y$. Then it follows from Lemma~\\ref{pi2} that $x$, $x'$, $y$ and $y'$ lie on the same minimal curve ordered in that way.\n\nAssume first that $ d(x_0,y_0) < d(x_0,y_1)$. We know from Lemma \\ref{pi2} that $d(x,y)\\leq d(x,y')$. On the other hand, we have\n\\begin{equation*}\n\\begin{split}\n d(x,y')&\\leq d(x,x_0) + d(x_0,y_0) + d(y_0,y')\\\\\n& \\leq d(x_0,y_0) + \\dfrac{\\delta}{2} + r \\\\\n&= d(x_0,y_1) + d(x_0,y_0) - d(x_0,y_1) + \\dfrac{\\delta}{2} + r \\\\\n&\\leq d(x_0,x) + d(x,y) + d(y,y_1) + d(x_0,y_0) - d(x_0,y_1) + \\dfrac{\\delta}{2} + r\\\\\n&\\leq d(x,y) + d(x_0,y_0) - d(x_0,y_1) + \\delta + 2r.\n\\end{split}\n\\end{equation*}\nIt follows that $d(x,y')< d(x,y)$ provided we take $r>0$ and $\\delta>0$ small enough which gives a contradiction. If $ d(x_0,y_0) = d(x_0,y_1)$, we have\n\\begin{equation*}\n\\begin{split}\nd(x_0,y_1) &\\leq d(x_0,x) + d(x,y) + d(y,y_1) \\\\\n& = d(x_0,x) + d(x,y') - d(y',y) + d(y,y_1)\\\\\n& \\leq d(x,y') - d(y',y) + \\dfrac{\\delta}{2} + r ,\n\\end{split}\n\\end{equation*}\n\\begin{equation*}\n d(x,y') \\leq d(x,x_0) + d(x_0,y_0) + d(y_0,y') \\leq d(x_0,y_0) + \\dfrac{\\delta}{2} + r,\n\\end{equation*}\n\\begin{equation*}\n d(y',y) \\geq d(y_0,y_1) - d(y_0,y') - d(y_1,y)\\geq d(y_0,y_1) -2r,\n\\end{equation*}\nhence,\n\\begin{equation*}\n d(x_0,y_1) \\leq d(x_0,y_0) - d(y_0,y_1) + 4r +\\delta. \n\\end{equation*}\nIt follows that $d(x_0,y_1) < d(x_0,y_0)$ provided we take $r>0$ and $\\delta>0$ small enough which gives also a contradiction.\n\\end{proof}\n\n\n\\section{Extension to more general metric measure spaces} \\label{extensions}\n\nFirst we note that a major part of intermediate steps in the strategy adopted in the present paper can be naturally extended to Polish and non-branching geodesic spaces equipped with a reference measure for which the Lebesgue's differentiation theorem holds. \n\nNext our choice of approximating costs $c_\\varepsilon$ in the approximation procedure is not the only possible one. This choice could in particular be adapted to fit other contexts (for instance concerning the relevant properties of solutions to the transport problem associated to the approximating cost).\n\nFinally the Measure Contraction Property is here technically very convenient. We note however that this property is unnecessarily too strong for what is actually needed in the proof about the lower density of the transport set. Much local and weaker versions about the behavior of the measure of sets transported along minimal curves are indeed sufficient as clearly shows up from the proof.\n\nThis approach can in particular be adapted to give an alternative proof of the existence of solutions to Monge's transport problem in the Riemannian setting without using Sudakov's type arguments.\n\nFor the reasons listed above it is furthermore very likely that the present strategy could be adapted and extended to other geodesic metric spaces.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\@ifstar{\\origsection*}{\\mysection}{\\@ifstar{\\origsection*}{\\mysection}}\n\\def\\mysection{\\@startsection{section}{1}\\z@{.7\\linespacing\\@plus\\linespacing}{\n.5\\linespacing}{\\normalfont\\scshape\\centering\\S}}\n\\makeatother\n\n\\usepackage{comment}\n\\usepackage{amsmath,amssymb,amsthm}\n\\usepackage{mathrsfs}\n\\usepackage{dsfont}\n\\usepackage{mathabx}\\changenotsign\n\\usepackage[ruled,linesnumbered,vlined]{algorithm2e}\n\n\\usepackage{tikz}\n\\usetikzlibrary{shapes,decorations,arrows,calc,arrows.meta,fit,positioning}\n\\usepackage{graphicx}\n\\usepackage{fullpage}\n\\usetikzlibrary{decorations.pathreplacing, matrix}\n\n\\usepackage{lineno}\n\n\\usepackage{microtype}\n\\usepackage{color}\n\\usepackage[utf8]{inputenc}\n\\usepackage{enumitem}\n\\newcommand\\rmlabel{\\upshape({\\itshape\\roman*\\,\\\/})}\n\\newcommand\\RMlabel{\\upshape(\\Roman*)}\n\\newcommand\\alabel{\\upshape({\\itshape\\alph*\\,\\\/})}\n\\newcommand\\Alabel{\\upshape({\\itshape\\Alph*\\,\\\/})}\n\\newcommand\\nlabel{\\upshape({\\itshape\\arabic*\\,\\\/})}\n\\newcommand\\nlabelp{\\upshape(P\\arabic*)}\n\n\\usepackage{xcolor}\n\\usepackage{hyperref}\n\\hypersetup{%\n colorlinks,\n linkcolor={red!60!black},\n citecolor={green!60!black},\n urlcolor={blue!60!black}\n}\n\n\\usepackage{tikz}\n\\usepackage{subcaption}\n\\usetikzlibrary{graphs}\n\n\\usepackage{bookmark}\n\n\\usepackage[T1]{fontenc}\n\\usepackage{lmodern}\n\n\\usepackage[english]{babel}\n\n\\usepackage{fullpage}\n\\usepackage{setspace}\n\\footskip28pt\n\n\\usepackage{enumitem}\n\n\\def\\itm#1{\\rm ({#1})}\n\\def\\itmit#1{\\itm{\\it #1\\,}}\n\\def\\itmit{\\roman{*}}{\\itmit{\\roman{*}}}\n\n\\newcommand{\\hat{r}}{\\hat{r}}\n\\let\\eps=\\varepsilon\n\\let\\polishlcross=\\ifmmode\\ell\\else\\polishlcross\\fi\n\\def\\ifmmode\\ell\\else\\polishlcross\\fi{\\ifmmode\\ell\\else\\polishlcross\\fi}\n\n\\newcommand\\tand{\\ \\text{and}\\ }\n\\newcommand\\qand{\\quad\\text{and}\\quad}\n\\newcommand\\qqand{\\qquad\\text{and}\\qquad}\n\n\\let\\emptyset=\\varnothing\n\\let\\setminus=\\smallsetminus\n\\let\\backslash=\\smallsetminus\n\n\\newcommand{\\mathbb{E}}{\\mathbb{E}}\n\\newcommand{\\mathbb{N}}{\\mathbb{N}}\n\\newcommand{\\mathbb{P}}{\\mathbb{P}}\n\n\\makeatletter\n\\def\\mathpalette\\mov@rlay{\\mathpalette\\mov@rlay}\n\\def\\mov@rlay#1#2{\\leavevmode\\vtop{%\n \\baselineskip\\z@skip \\lineskiplimit-\\maxdimen\n \\ialign{\\hfil$\\m@th#1##$\\hfil\\cr#2\\crcr}}}\n\\newcommand{\\charfusion}[3][\\mathord]{\n #1{\\ifx#1\\mathop\\vphantom{#2}\\fi\n \\mathpalette\\mov@rlay{#2\\cr#3}\n }\n \\ifx#1\\mathop\\expandafter\\displaylimits\\fi}\n\\makeatother\n\n\\newcommand{\\charfusion[\\mathbin]{\\cup}{\\cdot}}{\\charfusion[\\mathbin]{\\cup}{\\cdot}}\n\\newcommand{\\charfusion[\\mathop]{\\bigcup}{\\cdot}}{\\charfusion[\\mathop]{\\bigcup}{\\cdot}}\n\\newcommand{\\abs}[1]{\\left| #1 \\right|}\n\\newcommand{\\cbc}[1]{\\left\\lbrace #1 \\right\\rbrace}\n\\newcommand{\\bc}[1]{\\left( #1 \\right)}\n\\newcommand{\\floor}[1]{\\left\\lfloor #1 \\right\\rfloor}\n\\newcommand{\\ceil}[1]{\\left\\lceil #1 \\right\\rceil}\n\\newcommand{\\mhk}[1]{\\textcolor{red}{#1}}\n\\newcommand{\\op}[1]{\\textcolor{blue}{#1}}\n\n\n\\newtheorem{theorem}{Theorem}\n\\newtheorem{lemma}[theorem]{Lemma}\n\\newtheorem{corollary}[theorem]{Corollary}\n\\newtheorem{conjecture}[theorem]{Conjecture}\n\\newtheorem{problem}[theorem]{Problem}\n\\newtheorem{question}[theorem]{Question}\n\\newtheorem{property}[theorem]{Property}\n\\newtheorem{claim}[theorem]{Claim}\n\\newtheorem{subclaim}[theorem]{Sub-Claim}\n\\newtheorem{proposition}[theorem]{Proposition}\n\\newtheorem{definition}[theorem]{Definition}\n\\newtheorem{fact}[theorem]{Fact}\n\\newtheorem*{observation}{Observation}\n\n\\def\\mcarrow{\\xrightarrow[{\\raisebox{.5mm}[1mm][0mm]{$\\scriptstyle \\rm\np$}}]{\\raisebox{0.0mm}[0mm]{$\\scriptstyle \\rm rb$}}}\n\n\n\\def\\pmc#1{p^{\\rm rb}_{#1}}\n\\def\\mathcal{P}{\\mathcal{P}}\n\\def\\mathcal{W}{\\mathcal{W}}\n\\def\\mathcal{Z}{\\mathcal{Z}}\n\\def\\mathcal{A}{\\mathcal{A}}\n\\def\\mathcal{B}{\\mathcal{B}}\n\n\\usepackage{datetime}\n\\usepackage{lineno}\n\\newcommand*\\patchAmsMathEnvironmentForLineno[1]{%\n\\expandafter\\let\\csname old#1\\expandafter\\endcsname\\csname #1\\endcsname\n\\expandafter\\let\\csname oldend#1\\expandafter\\endcsname\\csname end#1\\endcsname\n\\renewenvironment{#1}%\n{\\linenomath\\csname old#1\\endcsname}%\n{\\csname oldend#1\\endcsname\\endlinenomath}\n\\newcommand*\\patchBothAmsMathEnvironmentsForLineno[1]{%\n\\patchAmsMathEnvironmentForLineno{#1}%\n\\patchAmsMathEnvironmentForLineno{#1*}}%\n\\AtBeginDocument{%\n\\patchBothAmsMathEnvironmentsForLineno{equation}%\n\\patchBothAmsMathEnvironmentsForLineno{align}%\n\\patchBothAmsMathEnvironmentsForLineno{flalign}%\n\\patchBothAmsMathEnvironmentsForLineno{alignat}%\n\\patchBothAmsMathEnvironmentsForLineno{gather}%\n\\patchBothAmsMathEnvironmentsForLineno{multline}%\n}\n\n\\newcommand{}{}\n\\def\\hfill\\scalebox{.6}{$\\Box$}{\\hfill\\scalebox{.6}{$\\Box$}}\n\n\\newenvironment{claimproof}[1][Proof]{\n \\renewcommand{}{\\qedsymbol}\n \\renewcommand{\\qedsymbol}{\\hfill\\scalebox{.6}{$\\Box$}}\n \\begin{proof}[#1]\n}{\n \\end{proof}\n \\renewcommand{\\qedsymbol}{}\n}\n\n\\newcommand{\\note}[1]{\\textcolor{red}{#1}}\n\n\\begin{document}\n\\onehalfspace\n\\shortdate\n\\yyyymmdddate\n\\settimeformat{ampmtime}\n\\date{\\today, \\currenttime}\n\n\\title{Minimum degree conditions for containing an $r$-regular $r$-connected subgraph}\n\n\\author[M.~Hahn-Klimroth]{Max Hahn-Klimroth} \n\\address{hahnklim@math.uni-frankfurt.de, Goethe University Frankfurt, Robert-Mayer-Str. 10, 60235 Frankfurt, Germany }\n\\thanks{MHK is supported by DFG grant CO 646\/5.}\n\n\\author[O.~Parczyk]{Olaf Parczyk}\n\\address{parczyk@mi.fu-berlin.de, FU Berlin, Arnimallee 3, 14195 Berlin, Germany}\n\\thanks{OP is supported by DFG grant PA 3513\/1-1.}\n\n\n\\author[Y.~Person]{Yury Person}\n\\address{yury.person@tu-ilmenau.de, TU Ilmenau, Weimarer Str. 25, 98684 Ilmenau, Germany} \n\\thanks{YP is supported by the Carl Zeiss Foundation and by DFG grant PE 2299\/3-1.}\n\n\\begin{abstract}\nWe study optimal minimum degree conditions when an $n$-vertex graph $G$ contains an $r$-regular $r$-connected subgraph. \nWe prove for $r$ fixed and $n$ large the condition to be $\\delta(G) \\ge \\frac{n+r-2}{2}$ when $nr \\equiv 0 \\pmod 2$. This answers a question of M.~Kriesell.\n\\end{abstract}\n\n\n\\maketitle\n\n\n\n\\@ifstar{\\origsection*}{\\mysection}{Introduction}\n\nA typical question in extremal graph theory is to determine (asymptotically) optimal minimum degree conditions for a graph $G$ on $n$ vertices to contain a given copy of some spanning graph. \nThe classical theorem of Dirac~\\cite{dirac} asserts the optimal minimum degree condition to contain a Hamilton cycle to be $\\tfrac n2$. There are numerous generalisations of this result to higher connected cycles (powers of Hamilton cycles)~\\cite{KSS_Seynmour}, which in turn generalise the theorems of Corradi and Hajnal~\\cite{CH63} and Hajnal and Szemer\\'edi~\\cite{HS_erdos} about clique factors in graphs. The most comprehensive result which asymptotically subsumes all of the mentioned results is the bandwidth theorem of B\\\"ottcher, Schacht and Taraz~\\cite{BST09}. This theorem provides a sufficient condition, which asymptotically depends only on the chromatic number of a bounded degree graph with sublinear bandwidth to be contained in a given dense graph. We also refer to the excellent survey~\\cite{KO09} by K\\\"uhn and Osthus for more results.\n\nThe present work is motivated by a question of Matthias Kriesell~\\cite{MKcomm} about optimal minimum degree condition sufficient to assert the existence of a $4$-regular $4$-connected spanning subgraph. This question in turn was motivated by the work of Bang-Jensen and Kriesell on good acyclic orientations of $4$-regular $4$-connected graphs~\\cite{BJK19}.\n\nWe answer Kriesell's question by proving the following general result about $r$-connected $r$-regular subgraphs of $G$.\n\\begin{theorem}\n \\label{thm:main}\n For any $r \\ge 2$ there exists an $n_0$ such that any $n$-vertex graph $G$ with minimum degree $\\delta(G) \\ge \\frac{n+r-2}{2}$, $n \\ge n_0$, and $nr \\equiv 0 \\pmod 2$ contains a spanning $r$-regular $r$-connected subgraph. \n\\end{theorem}\nNote that for $r \\ge 2$ an $n$-vertex graph $G$ with minimum degree $\\delta(G) \\ge \\frac{n+r-2}{2}$ always is $r$-connected, whereas one can easily come up with examples certifying the optimality of this result (e.g.\\ two $K_{(n+r)\/2}$'s sharing $r$ vertices).\nThe theorem above asserts that there are minimal $r$-connected subgraphs of $G$ which are in fact $r$-regular.\nObserve that for $r=2$ this follows immediately from Dirac's theorem~\\cite{dirac} with $n_0=3$, as a Hamilton cycle is $2$-regular and $2$-connected.\nOwing to the use of the regularity lemma the $n_0$ given by Theorem~\\ref{thm:main} will be very large.\n\nIn the following we briefly introduce some notation and discuss possible candidates for $r$-regular $r$-connected subgraphs that will be found in $G$ by Theorem~\\ref{thm:main}. \nThe \\emph{$t$-blow-up of a graph $F$} is obtained by replacing every vertex by $t$ vertices and every edge by a complete bipartite graph $K_{t,t}$.\nLet $C_n$ be the cycle on $n$ vertices and $P_n$ the $n$-vertex path.\nWe denote by $C_n(t)$ and $P_n(t)$ the $t$-blow-up of $C_n$ and $P_n$, respectively.\nWe use a similar definition for odd values of $t$.\nWe denote by $C_n(t-\\tfrac12)$ the $t$-blow-up of $C_n$ for $n$ even, where every other edge only gets a $K_{t,t}$ minus a perfect matching.\nSimilarly, $P_n(t-\\tfrac12)$ is the $t$-blow-up of the $n$-vertex path, where every other edge (starting with the first) only gets a $K_{t,t}$ minus a perfect matching.\nWe also call these the \\emph{$(t-\\tfrac 12)$-blow-ups}.\nNote that $C_n(t)$ is $2t$-regular and $C_n(t-\\tfrac 12)$ is $(2t-1)$-regular.\n\nIn most cases in our proof of Theorem~\\ref{thm:main} we will be able to find a spanning copy of an $\\tfrac r2$-blow-up of a cycle, while allowing other structures with all but a small fraction of vertices in $\\tfrac r2$-blow-ups of paths (see Section~\\ref{sec:o_constructions} for more details).\nHowever, when $n$ is even and not divisible by $4$, the graph $G$ obtained by taking the disjoint union of two cliques $K_{n\/2-2}$ and adding four additional vertices that are connected to all previous $n-4$ vertices cannot contain a copy of $C_n(4)$. \nFinally, observe, that the bandwidth theorem~\\cite{BST09} guarantees the asymptotically best minimum degree condition $\\tfrac n2 +o(n)$. Thus, Theorem~\\ref{thm:main} improves this asymptotic bound to the exact one.\n\nBeyond these blow-ups it would be interesting to study the minimum degree threshold for other spanning structures that can be obtained by identifying vertices or edges of copies of a small graph on a cycle.\nIn particular, when the small graph is not bipartite, this threshold can depend on its chromatic number or critical chromatic number similarly as when taking disjoint copies (see~\\cite{KO09}).\n\n\n\\subsection{Organisation of the paper}\nThe paper is structured as follows. In Section~\\ref{sec:tools} we collect the essential tools (regularity and blow-up lemmas), while Section~\\ref{sec:overview} provides a proof overview, which consists of three cases (extremal case I, extremal case II and non-extremal case). These cases are dealt with in the subsequent Sections~\\ref{sec:extremal1},~\\ref{sec:extremal2} and~\\ref{sec:non-extremal}.\n\n\\@ifstar{\\origsection*}{\\mysection}{Tools and Notation}\\label{sec:tools}\nFor standard graph theoretic definitions we refer to Bollob\\'as~\\cite{Bolbook98}. \nThe main tools are Szemer\u00e9di's regularity lemma~\\cite{Sze_regularity} and the blow-up lemma by Koml\u00f3s, S\u00e1rk\u00f6zy, and Szemer\u00e9di~\\cite{KSS_Blowup}.\nFor this let $G=(V,E)$ be a graph.\nFor any two sets $A,B \\subseteq V$ we denote by $e_G(A,B)$ the number of edges of $G$ with one endpoint in $A$ and one in $B$.\nThen the \\emph{density} $d(A,B)$ between these sets is $\\frac{e(A,B)}{|A||B|}$.\n\n\\begin{definition}\n The pair $(A,B)$ is \\emph{$\\varepsilon$-regular} if for all $X \\subseteq A$, $Y \\subseteq B$ with $|X| \\ge \\varepsilon |A|$, $|Y| \\ge \\varepsilon |B|$ we have $|d(X,Y)-d(A,B)| \\le \\varepsilon$. \n\\end{definition}\n\nThe following lemma guarantees that (not too small) induced subgraphs of $\\varepsilon$-regular pairs are still regular (although with a slightly worse parameter).\n\\begin{lemma}[Slicing lemma]\n\\label{lem:slicing}\nLet $(A,B)$ be an $\\eps$-regular pair with $d(A,B)=d$, let $\\tfrac 12 \\ge \\gamma > \\eps$, and $A' \\subseteq A$ and $B' \\subseteq B$ be of size $|A'| \\ge \\gamma |A|$ and $|B'|\\ge \\gamma |B|$.\nThen $(A',B')$ is $2\\eps$-regular pair with $d(A,B) \\ge d'$, where $|d-d'|\\le \\eps$.\\qed\n\\end{lemma}\n\nWhen working with the regular pairs, one often needs a somewhat stronger concept of super-regularity.\n\\begin{definition}\n The pair $(A,B)$ is an \\emph{$(\\varepsilon,\\delta)$-super-regular} pair if it is $\\varepsilon$-regular and $\\deg(a,B) \\ge \\delta |B|$, $\\deg(b,A) \\ge \\delta |A|$ for all $a \\in A$, $b \\in B$.\n\\end{definition}\n\nThe next lemma asserts that there every $\\varepsilon$-regular pair contains an almost spanning super-regular pair.\n\\begin{lemma}\n\\label{lem:superreg}\nLet $(A,B)$ be an $\\eps$-regular pair with $d(A,B)=d$.\nThen there exists $A'\\subseteq A$ and $B' \\subseteq B$ with $|A'|\\ge (1-\\eps) |A|$ and $|B'| \\ge (1-\\eps) |B|$ such that $(A',B')$ is a $(2\\eps,d-3\\eps)$-super-regular pair. \\qed\n\\end{lemma}\n\nWe will use the following degree form of the regularity lemma by Koml\u00f3s and Simonovits~\\cite{koml_simon}.\n\n\\begin{lemma}[Regularity lemma, degree version]\n \\label{lem:regularity}\n For every $\\varepsilon>0$ there exists an integer $M$ such that for any graph $G$ and $d \\in [0,1]$ there is a partition of $V(G)$ into $\\ell+1 \\le M$ clusters $V_0,\\dots,V_{\\ell}$ and a subgraph $G'$ of $G$ such that\n \\begin{enumerate}[label=\\upshape(P\\arabic*)]\n \\item \\label{reg:size} $|V_0| \\le \\varepsilon |V(G)|$ and $|V_i| = L \\le \\varepsilon |V(G)|$ for all $1 \\le i \\le \\ell$.\n \\item \\label{reg:deg} $\\deg_{G'}(v) \\ge \\deg_G(v) - (d+\\varepsilon) |V|$ for all $v \\in V$.\n \\item \\label{reg:ind} For $1 \\le i \\le \\ell$ the set $V_i$ is independent in $G'$.\n \\item \\label{reg:reg} For $1 \\le i < j \\le \\ell$ the pair $(V_i,V_j)$ is $\\varepsilon$-regular in $G'$ and has density $0$ or $d$.\n \\end{enumerate}\n\\end{lemma}\n\nThe blow-up lemma allows us to embed spanning subgraphs with bounded degree. We will use the following special case deduced from~\\cite[Remark~13]{KSS_Blowup}.\n\n\\begin{lemma}[Bipartite blow-up lemma]\n \\label{lem:blowup}\n For each $d,c>0$ and integer $\\Delta$ there exist $\\varepsilon>0$, $\\alpha>0$ and integer $n_0$ such that the following holds for any $n \\ge n_0$.\n Let $H$ be a bipartite graphs on classes $A$ and $B$ with $|A|=|B|=n$ such that $(A,B)$ is a $(\\varepsilon,d)$-super-regular pair and let $G$ be a bipartite graph on classes $X$ and $Y$ with $|X|=|Y|=n$ that has maximum degree bounded by $\\Delta$.\n Moreover, for any $X' \\subseteq X$ and $Y' \\subseteq Y$ with $|X'|,|Y'| \\le \\eps n$ let $A_x \\subseteq A$ and $B_y \\subseteq B$ for each $x \\in X'$ and $y \\in Y'$ with $|A_x|,|B_x| \\ge cn$.\n Then there exists an embedding of $G$ into $H$ such that all $x \\in X'$ and $y \\in Y'$ are embedded into $A_x$ and $B_y$, respectively.\n\\end{lemma}\n\nWe remark that in our application $X'$ and $Y'$ will be of constant size and all $A_x$ and all $B_y$ will be the same.\n\n\\@ifstar{\\origsection*}{\\mysection}{Proof overview}\\label{sec:overview}\n\nThe proof of Theorem~\\ref{thm:main} will be split into three cases.\nWe now explain this case distinction and then give an overview of the proof for each of these cases.\nLet $G$ be a graph with minimum degree $\\tfrac{n+r-2}{2}$.\nFor $\\alpha>0$ we call $G$ \\emph{$\\alpha$-extremal} if there are two sets $A,B \\subseteq V(G)$ of size $(\\frac{1}{2}-\\alpha)n \\le |A|,|B| \\le \\frac{n}{2}$ such that $d(A,B) < \\alpha$.\nWith the help of the regularity lemma we will cover the case that $G$ is not $\\alpha$-extremal for any $\\tfrac{1}{32}>\\alpha>0$ in Section~\\ref{sec:non-extremal}.\n\nSo we can assume that $G$ is $\\alpha$-extremal for some $\\alpha>0$.\nUsing the minimum degree condition in $G$ it is easy to see that the sets $A$ and $B$ have to be almost disjoint or almost the same.\nThis implies that $G$ contains a large set that is 'almost' independent or it is 'close' to the disjoint union of two cliques $K_{n\/2}$.\nMore precisely, there exists $\\alpha'>0$ such that one of the following holds:\nEither, there are two disjoint sets $A,B \\subseteq V(G)$ with $(\\tfrac 12 -\\alpha')n \\le |A|,|B| \\le (\\tfrac 12 +\\alpha')n$ such that $G[A]$ and $G[B]$ have minimum degree $(\\tfrac 12 -3\\alpha')n$ and every vertex outside of $A \\cup B$ has degree at least $\\alpha'n$ into $A$ and $B$ -- this will be the first extremal case treated in Section~\\ref{sec:extremal1}.\nOr, there is one set $A \\subseteq V(G)$ with $(\\tfrac12 -\\alpha')n \\le |A| \\le (\\tfrac12 +\\alpha')n$ is such that any vertex in $A$ has degree at least $(\\tfrac12 -3\\alpha')n$ into $V(G) \\setminus A$ and every vertex outside of $A$ has degree at least $3 \\alpha' n$ into $A$ -- this is the second extremal case treated in Section~\\ref{sec:extremal2}.\n\nTherefore, when choosing $0<\\alpha< \\tfrac{1}{32}$ sufficiently small for both extremal cases and the remaining cases will be `non-extremal'.\nThis implies Theorem~\\ref{thm:main}.\nIn the remainder of this section we sketch the argument for each of the three cases and afterwards explain why our constructions are indeed $r$-connected.\n\n\\subsection{Non-Extremal Case}\n\\label{sec:o_non-extremal}\n\nWe would like to find a spanning copy of $C_k(\\tfrac r2)$ in $G$, but an obvious necessary condition for this is that $v(G) \\equiv 0 \\pmod{2 \\lceil \\tfrac r2 \\rceil}$.\nIf this condition is satisfied, we will succeed, and, otherwise, find a slightly locally modified version.\nFor the proof we will have constants\n\\begin{align*}\n \\varepsilon \\ll \\nu \\ll d \\ll \\beta \\ll \\alpha < \\frac{1}{32}\n\\end{align*}\nand $s = \\lceil \\tfrac r2 \\rceil$.\nWe follow similar arguments as in~\\cite{KSS_square}, which can be summarised by the following procedure:\n\\begin{enumerate}[label=\\upshape\\bf Step \\arabic*]\n \\item Apply regularity lemma (Lemma~\\ref{lem:regularity}) with $\\varepsilon$ and $d$ to obtain a regular partition of $G$.\\label{step:regularise}\n \\item Find $\\ell$ $\\varepsilon$-regular pairs $(X_i,Y_i)$ covering all but a small set $V_0$ with $|V_0| \\le 20 dn$.\\label{step:matching}\n \\item For $i=1,\\dots,\\ell$ connect $Y_i$ to $X_{i+1}$ with the $\\tfrac r2$-blow-up of a path that we denote by $P_i$.\\label{step:connect}\n \\item For $i=1,\\dots,\\ell$ turn $(X_i,Y_i)$ into an $(\\varepsilon,d-\\varepsilon)$-super-regular pair with $|X_i|=|Y_i|$, slightly increasing $V_0$ to $|V_0| \\le 23 dn$.\\label{step:superreg}\n \\item Repeatedly take $\\nu n$ vertices from $V_0$ and append them to the paths $P_i$.\\label{step:absorb}\n \\item For $i=1,\\dots,\\ell$ use blow-up lemma (Lemma~\\ref{lem:blowup}) to find a spanning copy of an $\\tfrac r2$-blow-up of a path in $(X_i,Y_i)$ connecting $P_{i-1}$ with $P_{i}$.\\label{step:spanning}\n\\end{enumerate}\n\nThe index $\\ell+1$ corresponds to $1$.\n\\ref{step:regularise} is natural and for~\\ref{step:matching} it is enough to find a large matching in a graph with minimum degree close to $\\frac{n}{2}$.\nDuring the performance of~\\ref{step:absorb} the degree of some vertices might get too small.\nIn this case we add them to a set $Q$ that we take care of before the next round.\nThis terminates as in every execution there are at most $3 \\varepsilon n \\ll \\nu n$ vertices added to $Q$.\nApart from this~\\ref{step:absorb} is very similar to~\\ref{step:connect}, which we now sketch with more details.\n\nLet $X$, $Y$ be the clusters that we want to connect with the $\\tfrac r2$-blow-up of a path $P$.\nIf there is a cluster $Z$ such that $(X,Z)$ and $(Z,Y)$ are $\\varepsilon$-regular pairs with density at least $d$ then we can easily find this path.\nOtherwise, let $A$ be the union of all clusters $Z$ such that $(X,Z)$ is an $\\varepsilon$-regular pair with density at least $d$ and $B$ the union of all clusters $Z$ for $(Y,Z)$ analogously.\nBy the minimum degree property in the cluster graph we get $|A|,|B| \\ge (\\frac{1}{2}-\\alpha)n$.\nAs $G$ is not $\\alpha$-extremal we have $d(A,B) > \\alpha$.\nTherefore, there exist two clusters $Z_1 \\in A$ and $Z_2 \\in B$ with $d(Z_1,Z_2) \\ge \\alpha$ and then $(X,Z_1)$, $(Z_1,Z_2)$, and $(Z_2,Y)$ are $\\varepsilon$-regular pairs with density at least $d$.\nThen it is again easy to find the path that we are interested in by following these three regular pairs.\n\nWe have to ensure that the end vertices of the paths always have high degree into the other cluster of the respective super-regular pair, because we want to connect them later and keep them through~\\ref{step:superreg}.\nFurthermore, we have to ensure that in~\\ref{step:absorb} the sizes of the $(\\varepsilon,d-\\varepsilon)$-super-regular pairs remain balanced.\nWe will give the details in Section~\\ref{sec:non-extremal}.\n\n\\subsection{Extremal Case I}\n\\label{sec:o_extremal1}\n\nIn this extremal case we will not use the regularity lemma, but the blow-up lemma will be helpful.\nRecall that in this case $G$ is `close' to the union of two disjoint cliques of size roughly $\\tfrac n2$ on vertex sets $A$ and $B$.\nThe main challenge is to find a bridge that connects both these cliques.\nIt is then easy to find the desired structure using the high degrees.\n\\begin{enumerate}[label=\\upshape\\bf Step \\arabic*]\n\\item \\label{stepd:bridge} In the case when $r$ is even the bridge will be a matching of size $r$ between $A$ and $B$ such that the end-vertices are well connected on their side.\nThe odd case is a little more delicate and we will find a matching of size $r+1$ or $r$ depending on the size of $V(G) \\setminus \\bc{A \\cup B}$ and the parity of $A$ and $B$.\n\\item \\label{stepd:absorb} Absorb all vertices that do not not belong to $A$ or $B$ by extending both ends of the path.\nWe can ensure that the left-over on each side has size divisible by $2r$.\n\\item \\label{stepd:cover} It is easy to see that the left-over on both sides can be split into a super-regular pair and that we can cover both with the $\\tfrac r2$-blow-up of a path using Lemma~\\ref{lem:blowup}.\n\\end{enumerate}\nIf we take care of the end-tuples between each of the steps this gives an $r$-regular $r$-connected path-structure covering $G$.\nIn Section~\\ref{sec:extremal1} we will give the details of the even and odd case separately.\n\n\\subsection{Extremal Case II}\n\\label{sec:o_extremal2}\n\nAgain, we will not use the regularity lemma in this part, but the blow-up lemma will still be helpful.\nWe can assume that we have a partition of $V(G)$ into $A$ and $B$ of size $(\\tfrac 12 \\pm \\alpha)n$ such that between these sets we have minimum degree $\\alpha n$ and all but at most $\\alpha n$ vertices from $A$ (or $B$) have degree $|B|-\\alpha n$ (or $|A|-\\alpha n$) into $B$ (or $A$).\nW.l.o.g.~assume that $|A|+m=\\tfrac 12 n=|B|-m$, where $0 \\le m \\le \\alpha n$.\nNote that in $G[B]$ we have minimum degree at least $m+\\tfrac{r-2}{2}$.\nLet $s=\\lceil \\tfrac r2 \\rceil$.\n\\begin{enumerate}[label=\\upshape\\bf Step \\arabic*]\n \\item \\label{stepe:find} If $\\Delta(G[B]) \\le 2 r \\alpha n$ find $m$ copies of $K_{1,s}$, such that all vertices are well connected to the other side. Otherwise, separate the vertices with higher degrees, then find copies of $K_{1,s}$, and afterwards find additional copies of $K_{1,r}$, such that the leaves are well connected.\n \\item \\label{stepe:absorb} Absorb these copies of $K_{1,s}$ and $K_{1,r}$ into an $r$-regular path-structure and then connect these together into one longer path-structure.\n After removing the path that we constructed we are left with sets $A_1\\subseteq A$ and $B_1 \\subseteq B$ with $|A_1|=|B_1|$.\n \\item \\label{stepe:absorb2} Absorb all vertices that do not have large degree to the other side into the path by alternating between both sides.\n After removing these vertices we are left with sets $A_2\\subseteq A_1$ and $B_2 \\subseteq B_1$ with $|A_2|=|B_2|$ and the property that all vertices have large degree to the other side.\n \\item \\label{stepe:cover} It is easy to see that $(A_2,B_2)$ is a super-regular pair and that we can cover it with the $\\tfrac r2$-blow-up of a path using Lemma~\\ref{lem:blowup}.\n\\end{enumerate}\nIf we take care of the end-tuples between each of the steps this gives an $r$-regular $r$-connected path-structure covering $G$.\nFor the first step we use the following.\n\\begin{lemma}\n\\label{lem:stars}\n For any integer $s$ there exists $\\alpha>0$ such that the following holds.\n Let $G$ be an $n$ vertex graph with maximum degree $\\Delta(G) \\le 4 s \\alpha n$ and minimum degree $\\delta(G) \\ge m + s - 1$, where $1 \\le m \\le \\alpha n$.\n Then there are $2m$ pairwise disjoint copies of $K_{1,s}$ in $G$.\n\\end{lemma}\n\nThe proof of this lemma and the second extremal case will be given in Section~\\ref{sec:extremal2}.\n\n\\subsection{Constructions}\n\\label{sec:o_constructions}\nFirst recall that the $\\tfrac r2$-blow-up of a cycle is $r$-regular and also $r$-connected.\nIt will not always be possible to construct this, but it will be the basic building block.\nWe might need to absorb some exceptional vertices, for example, when $n$ is not divisible by $r$.\nIn the case when $r$ is even we then remove a perfect matching from one $K_{s,s}$ and add one vertex that is connected to all $2s=r$ vertices that just lost one neighbour (c.f.~Figures~\\ref{fig:absorber_even},~\\ref{fig:absorbK1s}, and~\\ref{fig:Step54}).\nThe resulting graph is still $r$-connected, because we can not disconnect this part of the cycle by removing less than $\\tfrac r2$ vertices.\nA similar construction will be used in the case when $r$ is odd (c.f.~Figures~\\ref{fig:absorber_odd},~\\ref{fig:absorbK1s}~and~\\ref{fig:Step55}) that also preserves $r$-connectivity.\nApart from this, we also have to connect to $\\tfrac r2$-blow-ups of cycles by using at most $r$ edges between them (c.f.~Step~\\ref{stepd:bridge} of Section~\\ref{sec:o_extremal1}).\nWe will only need to take care of a small linear fraction of the vertices from $G$ and, therefore, almost all vertices are in the $\\tfrac r2$-blow-up of a path.\n\n\\@ifstar{\\origsection*}{\\mysection}{Extremal Case I}\n\\label{sec:extremal1}\nIn this section we deal with the first extremal case.\nWe will not use the regularity lemma in this part, but the blow-up lemma will still be helpful.\n\n\\begin{proof}[Proof of Extremal Case I]\nLet $r \\geq 3$ be an integer, let $\\eps>0$ be given by Lemma~\\ref{lem:blowup} on input $\\tfrac 12$, $\\tfrac 12$, and $r$ and let $0<\\alpha \\le \\eps (1000 r^2)^{-1}$.\nLet $G$ be an $n$-vertex graph with $\\delta(G) \\geq \\tfrac {n+r-2}{2}$ and let $A,B \\subseteq V(G)$ with $(\\tfrac 12-\\alpha)n \\le |A|,|B| \\le (\\tfrac 12+\\alpha)n$ such that $G[A]$ and $G[B]$ have minimum degree $(\\tfrac 12-3\\alpha)n$ and every vertex in $C=V(G) \\setminus (A \\cup B)$ has degree at least $\\alpha n$ into $A$ and $B$.\nOur goal is to find an $r$-regular, $r$-connected spanning subgraph in $G$ provided that $n$ is large enough. \n\n\\subsection{The even case}\n\nAssume that $r$ is even.\nWe begin by constructing $\\tfrac r2$ bridges of size 2 between $A$ and $B$ (\\ref{stepd:bridge} of Section \\ref{sec:o_extremal1}). A visualisation can be found in Figure~\\ref{fig:bridge_graph_h}.\n\n\n\\begin{figure}\n \\centering\n \\begin{tikzpicture}\n \\node[circle, draw=black] (a_11) at (0, 3) {$a_{11}$};\n \\node[circle, draw=black] (a_12) at (0, 1.5) {$a_{12}$};\n \\node[circle, draw=black] (b_11) at (0, -1.5) {$b_{11}$};\n \\node[circle, draw=black] (b_12) at (0, -3) {$b_{12}$};\n \n \\node[circle, draw=black] (a_21) at (2, 3) {$a_{21}$};\n \\node[circle, draw=green] (xa1) at (2, 1.5) {$x_{a_1}$};\n \\node[circle, draw=green] (xb1) at (2, -1.5) {$x_{b_1}$};\n \\node[circle, draw=black] (b_22) at (2, -3) {$b_{22}$};\n \n \\node[circle, draw=black] (a_31) at (4, 3) {$a_{31}$};\n \\node[circle, draw=green] (xa2) at (4, 1.5) {$x_{a_2}$};\n \\node[circle, draw=green] (xb2) at (4, -1.5) {$x_{b_2}$};\n \\node[circle, draw=black] (b_32) at (4, -3) {$b_{32}$};\n \n \\node[circle, draw=black] (a_41) at (6, 3) {$a_{41}$};\n \\node[circle, draw=black] (a_42) at (6, 1.5) {$a_{42}$};\n \\node[circle, draw=black] (b_41) at (6, -1.5) {$b_{41}$};\n \\node[circle, draw=black] (b_42) at (6, -3) {$b_{42}$};\n \n \n \\path[-] (a_11) edge[draw=black] (a_21);\n \\path[-] (a_12) edge[draw=black] (xa1);\n \\path[-] (a_11) edge[draw=black] (xa1);\n \\path[-] (a_12) edge[draw=black] (a_21);\n \n \\path[-] (xa1) edge[draw=black] (a_31);\n \\path[-] (a_21) edge[draw=black] (xa2);\n \\path[-] (a_21) edge[draw=black] (a_31);\n \n \\path[-] (a_31) edge[draw=black] (a_41);\n \\path[-] (xa2) edge[draw=black] (a_42);\n \\path[-] (a_31) edge[draw=black] (a_42);\n \\path[-] (xa2) edge[draw=black] (a_41);\n \n \n \\path[-] (b_11) edge[draw=black] (xb1);\n \\path[-] (b_12) edge[draw=black] (b_22);\n \\path[-] (b_11) edge[draw=black] (b_22);\n \\path[-] (b_12) edge[draw=black] (xb1);\n \n \\path[-] (b_22) edge[draw=black] (xb2);\n \\path[-] (xb1) edge[draw=black] (b_32);\n \\path[-] (b_22) edge[draw=black] (b_32);\n \n \\path[-] (xb2) edge[draw=black] (b_41);\n \\path[-] (b_32) edge[draw=black] (b_42);\n \\path[-] (xb2) edge[draw=black] (b_42);\n \\path[-] (b_32) edge[draw=black] (b_41);\n \n \\path[-] (xb2) edge[draw=green] (xa2);\n \\path[-] (xb1) edge[draw=green] (xa1);\n\n \\end{tikzpicture}\n \\caption{Bridge between the sets $A$ and $B$ in the special case $r = 4$.}\n \\label{fig:bridge_graph_h}\n\\end{figure}\n\n\n\\begin{claim}\\label{lemma_matching_edges}\nSuppose $\\delta(G) \\geq \\tfrac {n+r-2}{2}$ and $\\abs{A} \\leq \\abs{B}$. There is a matching $(x_{a_1}x_{b_1}, \\ldots, x_{a_r}x_{b_r})$ such that $\\abs{N(x_{a_i}) \\cap A} \\geq \\tfrac n5$ and $\\abs{N(x_{b_j}) \\cap B} \\geq \\tfrac n5$ for all $i, j \\leq r$.\n\\end{claim}\n\\begin{claimproof}\n In order to construct the matching, it suffices to find $r$ edges from $A$ to $V \\setminus A$. Indeed, suppose we find the $r$ edges $a_1c_1, \\ldots, a_rc_r$. If $c_i \\in B$, we take this edge. If $c_i \\in V \\setminus (A \\cup B)$, $c_i$ has either $\\tfrac n5$ edges into $B$ (in this case, take edge $a_ic_i$), or it has $\\tfrac n5$ edges into $A$. Let $i_1, \\ldots, i_l$ be the indices such that $c_{i_j}$ does not have $\\tfrac n5$ edges into $B$. By definition of $\\alpha$-extremity, each $c_{i_j}$ has $\\alpha n$ neighbors in $B$. Select $b_{i_1}, \\ldots, b_{i_l}$ s.t. $b_{i_j} \\in N(c_{i_j}) \\cap B$ and $b_{i_j} \\neq b_{i_k}$ for all $j \\neq k$ (and being disjoint from those $c_i \\in B$ (this is clearly possible). We add edges $b_{i_j}c_{i_j}$ to the matching. \n \nIt remains to show that these edges exist. \nFirst, suppose that $n$ is even. If $\\abs{A} \\leq \\tfrac{n-r}{2}$, the minimum degree of $ \\tfrac{n+r-2}{2}$ guarantees that each vertex of $A$ needs to find at least $\\tfrac{n+r-2}{2} - (\\tfrac{n-r}{2} - 1) = r$ neighbors outside of $A$, hence the assertion follows.\nSuppose $\\abs{A} = \\tfrac{n-r}{2} + i$ with $i = 1, \\ldots, \\tfrac r2$. In this case, $\\abs{V \\setminus A} = \\tfrac{n+r}{2} - i$. Each vertex of $A$ finds at least $r-i$ neighbors outside of $A$. Suppose that $N(A) \\setminus A$ has size at most $r-1$ (thus, all edges from $A$ into the rest of the graph belong to $r-1$ vertices). Now pick a different vertex in the complement (which exists, as $\\tfrac{n+r}{2} - i \\gg r$). This vertex requires $\\tfrac{n+r-2}{2} - (\\tfrac{n+r}{2} - i - 1) = i$ neighbors in $A$, which is a contradiction.\n\nNow, if $n$ is odd, because the minimum degree needs to be an integer, it is at least $\\tfrac{n+1+r-2}{2}$, hence upon removal of one vertex, we are left with a graph on $n' = n-1$ vertices and minimum degree at least $\\tfrac{n+1+r-2}{2}-1 = \\tfrac{n'+r-2}{2}$ (and $n'$ being even). Hence the assertion follows from the previous discussion.\n\\end{claimproof}\n\nTherefore, Claim~\\ref{lemma_matching_edges} gives us the green sub-structure of Figure \\ref{fig:bridge_graph_h}. \nNow, we take two of those matching edges (think of them as being $\\tfrac r2$ pairs of $2$ edges). Denote the vertices that are connected to at least $\\tfrac n5$ vertices in $A$ as $x_{a_1}, x_{a_2}$. We next prove that the black structure around $x_{a_1}, x_{a_2}$ shown Figure \\ref{fig:bridge_graph_h} exists.\n\n\\begin{claim}\\label{bridge_graph_existence}\nThere are distinct vertices $a_{i,1}, \\ldots, a_{i, r\/2} \\in A$ for $i=1,4$ and $a_{i,1}, \\ldots, a_{i, r\/2-1} \\in A$ for $i=2,3$ with the following properties.\n\\begin{enumerate}\n \\item The edges $a_{i, j}a_{i+1, k}$ for $i=1, 2 ,3$ and $j = 1 , \\ldots, \\tfrac r2$ (or $\\tfrac r2 - 1$, respectively) exist,\n \\item the edges $x_{a_1}a_{1, j}$ and $x_{a_2}a_{4, j}$ exist for $j = 1 , \\ldots,\\tfrac r2$,\n \\item the edges $x_{a_1}a_{3, j}$ and $x_{a_2}a_{1, j}$ exist for $j = 1 , \\ldots,\\tfrac r2 -1$.\n\\end{enumerate}\n\\end{claim}\n\n\\begin{claimproof}\nWe select $r - 1$ vertices $a_{1,i}$, $a_{3, j} \\in N(x_{a_1}) \\cap A$ arbitrarily (but disjoint from $x_{a_2}$). Those exist as $x_{a_1}$ has at least $\\tfrac n5$ neighbors in $A$. Each of those vertices is connected to at least $(\\tfrac 12 - 3 \\alpha)n$ vertices in $A$, hence each vertex has at least $(\\tfrac 12 - 3 \\alpha)n - r - 1$ neighbors in $A$ that do not belong to $a_{1,i}$, $a_{3, j}$ or $x_{a_1}, x_{a_2}$.\nTherefore, the joint neighborhood \n\\[N := \\bc{A \\cap \\bigcap_{i = 1}^{r\/2} N(a_{1,i}) \\bigcap_{j=1}^{r\/2 - 1} N(a_{1,j})} \\setminus \\bc{\\bigcup_{i=1}^{r\/2} \\cbc{ a_{4, i} } \\bigcup_{j=1}^{r\/2-1} a_{3, j} \\cup \\cbc{ x_{a_1}, x_{a_2} }}\\]\nhas size at least $(\\tfrac 12 - 4 r \\alpha)n$. Therefore, we find \n\\[\\abs{ N(x_{a_2}) \\cap N } \\geq \\frac{n}{100} \\, , \\]\nthus the claim follows as the same token holds in $B$ as well.\n\\end{claimproof}\nWe denote the resulting collection of vertices in $A$ by $X_{A, 1}, \\ldots, X_{A, r\/2}$.\nClearly, each vertex in $A$ stays connected to at least $(\\tfrac 12 - 4 \\alpha)n$ vertices in\n\\[A'_{0} = A \\setminus \\bc{ V(X_{A, 1}) \\cup \\ldots \\cup V(X_{A, r\/2}) } \\, .\\]\n\nNext, we introduce a \\textit{gluing operation} GE.\n\\begin{claim}[Gluing operation GE]\nGiven two disjoint sets $D_1, D_2 \\subset A'_0$ of size exactly $\\tfrac r2$, we find two disjoint sets $D, D' \\subset A'_0 \\setminus (D_1 \\cup D_2)$ of size $\\tfrac r2$ such that\n\\[G[D_1, D] \\equiv K_{r\/2, r\/2}, \\quad G[D, D'] \\equiv K_{r\/2, r\/2}\\quad \\text{and} \\quad G[D', D_2] \\equiv K_{r\/2, r\/2} \\, .\\]\n\\end{claim}\n\\begin{claimproof}\nAs the joint $A'_0$ - neighborhood of $D_1$ and $D_2$ has size at least $(\\tfrac 12 - 10r \\alpha)n$, the assertion follows.\n\\end{claimproof}\n\nUsing GE, we glue the $\\tfrac r2$ bridges in $A$ and $B$ respectively together using mutually disjoint vertex sets $D^A_1, \\ldots,D^A_{r\/2 - 1}$ and $D^B_1, \\ldots,D^B_{r\/2 - 1}$ and are left with path-like structures $P_A$ and $P_B$. After gluing the bridges together, we let $A' = A'_0 \\setminus V(P_A), B' = B'_0 \\setminus V(P_B)$. \n\nIn a next step we need to absorb left-over vertices (\\ref{stepd:absorb} of Section \\ref{sec:o_extremal1}). To this end define two absorber-graphs for a vertex $u$: $\\xi_r(u)$ and $\\xi'_r(u)$ (see Figure \\ref{fig:absorber_even}).\n\\begin{definition}\n Let $D \\in \\cbc{A', B'}$ and $u$ a vertex such that $\\abs{ N(u) \\cap D } \\geq \\tfrac n6$. Define $\\xi_r(u)$ as follows.\n \\begin{itemize}\n \\item Select $D_1 = \\cbc{d_1, d_2, \\ldots, d_{r\/2}}, D_2 = \\cbc{d'_1, \\ldots, d'_{r\/2}} \\subset N(u) \\cap D$, hence $r$ pairwise disjoint vertices.\n \\item Select $D' = \\cbc{ u'_1, \\ldots, u'_{r\/2 - 1}} \\subset N(D_1) \\cap N(D_2) \\cap D \\setminus (D_1 \\cup D_2 \\cup \\cbc{u}) $.\n \\end{itemize}\n Define $\\xi_r(u)$ as the graph containing $D_1, D_2, D'$ and $u$ as well as all the edges from $D_1$ to $D' \\cup \\cbc{u}$ and from $D_2$ to $D' \\cup \\cbc{u}$.\n Furthermore, define $\\xi'_r(u)$ via\n \\begin{itemize}\n \\item Select $D_1 = \\cbc{d_1, d_2, \\ldots, d_{r\/2}}, D_2 = \\cbc{d'_1, \\ldots, d'_{r\/2}} \\subset N(u) \\cap D$, hence $r$ pairwise disjoint vertices.\n \\item Select $D' = \\cbc{ u'_1, \\ldots, u'_{r\/2 - 1}} \\subset N(D_1) \\cap N(D_2) \\cap D \\setminus (D_1 \\cup D_2 \\cup \\cbc{u}) $.\n \\item Select $\\tfrac r2$ vertices $E_0 = \\cbc{e_0, \\ldots, e_{r\/2-1}}$ and an additional disjoint vertex $e_{r\/2}$ from $D \\setminus \\bc{ D_1 \\cup D_2 \\cup D' }$ such that $E_0 \\cup \\cbc{e_2} \\subset N(d_2) \\cap \\ldots \\cap N(d_{r\/2}) \\cap D$ and $G[E_0, \\cbc{e_{r\/2}}] \\equiv K_{r\/2, 1}$. \n \\end{itemize}\n Define $\\xi'_r(u)$ as the graph containing $D_1, D_2, D'$ and $u$ as well as all the edges from $D_1$ to $D' \\cup \\cbc{u}$ and from $D_2$ to $D' \\cup \\cbc{u}$.\n\\end{definition}\nClearly, by the sizes of $A', B'$ and the minimum-degree condition, given at most $100 \\alpha n$ pairwise different vertices, there is an absorber for each vertex which is disjoint from all other absorbers. Furthermore, by the minimum degree condition inside of $A'$ and $B'$, this family of absorbers exists.\n\\begin{figure}\n \\centering\n \\begin{tikzpicture}\n \n\n \\node[circle, draw=green!40] (d1) at (2, 1) {$d_1$};\n \\node[circle, draw=green!40] (d2) at (2, -1) {$d_2$};\n\n \\node[circle, draw=green!40] (dp1) at (6, 1) {$d'_{1}$};\n \\node[circle, draw=green!40] (dp2) at (6, -1) {$d'_{2}$};\n \n \\node[circle, draw=blue!40] (up1) at (4, 1) {$u'_{1}$};\n \\node[circle, draw=green] (u) at (4, -1) {$u$};\n\n \n \\path[-] (u) edge[draw=green] (d1);\n \\path[-] (u) edge[draw=green] (d2);\n \\path[-] (u) edge[draw=green] (dp1);\n \\path[-] (u) edge[draw=green] (dp2);\n \n \\path[-] (d1) edge[draw=blue] (up1);\n \\path[-] (d2) edge[draw=blue] (up1);\n \\path[-] (dp1) edge[draw=blue] (up1);\n \\path[-] (dp2) edge[draw=blue] (up1);\n \n \n \n \\node[circle, draw=blue!40] (e0) at (10, 1) {$e_0$};\n \\node[circle, draw=blue!40] (e1) at (10, -1) {$e_1$};\n \\node[circle, draw=blue!40] (e2) at (11, 0) {$e_2$};\n \n \\node[circle, draw=green!40] (ld1) at (12, 1) {$d_1$};\n \\node[circle, draw=green!40] (ld2) at (12, -1) {$d_2$};\n\n \\node[circle, draw=green!40] (ldp1) at (16, 1) {$d'_{1}$};\n \\node[circle, draw=green!40] (ldp2) at (16, -1) {$d'_{2}$};\n \n \\node[circle, draw=blue!40] (lup1) at (14, 1) {$u'_{1}$};\n \\node[circle, draw=green] (lu) at (14, -1) {$u$};\n\n \n \\path[-] (lu) edge[draw=green] (ld1);\n \\path[-] (lu) edge[draw=green] (ld2);\n \\path[-] (lu) edge[draw=green] (ldp1);\n \\path[-] (lu) edge[draw=green] (ldp2);\n \n \\path[-] (ld1) edge[draw=blue] (lup1);\n \\path[-] (ld2) edge[draw=blue] (lup1);\n \\path[-] (ldp1) edge[draw=blue] (lup1);\n \\path[-] (ldp2) edge[draw=blue] (lup1);\n \n \\path[-] (ld1) edge[draw=blue] (e2);\n \\path[-] (ld2) edge[draw=blue] (e2);\n \\path[-] (e1) edge[draw=blue] (e2);\n \\path[-] (e0) edge[draw=blue] (e2);\n \n \\path[-] (e0) edge[draw=blue, bend left=30] (ld2);\n \\path[-] (e1) edge[draw=blue, bend left=30] (ld1);\n \n \\end{tikzpicture}\n \n \\caption{Absorbers $\\xi_4(u)$ (left) and $\\xi'_4(u)$ (right) with $r=4$, where $u$ is the green vertex, the green vertices are inside the $A'$-neighborhood of $u$ and the blue vertices are vertices chosen from $A'$.}\n \\label{fig:absorber_even}\n\\end{figure}\n\nWe are now in position to absorb the exceptional set $C$ (of course, without the bridging vertices on $P_A$ and $P_B$. For all up to $6 \\alpha n$ vertices in $C$ create a disjoint absorber $\\xi_r(u)$ as above. Chose as $D$ either $A'$ (if a vertex has $\\tfrac n5$ neighbors in $A$), or $B'$ otherwise. Next, using the gluing operation GE up to $6 \\alpha n$ times, glue the absorbers inside of $A'$ and $B'$ together, always using only vertices that did not get used in a previous gluing step or are part of the absorbers. As one only requires at most $24 r \\alpha n$ vertices during this procedure, it is clearly possible by above discussion.\n\nFinally, use GE again to glue the series of absorbers to $P_A$ and $P_B$ respectively (which is clearly possible, as this is only one operation on each set).\n\nAfter the repetitive gluing, we are left with sets $A''$ and $B''$ (hence, $A'$ without the glued structures $P'_A$) and the path-like subgraph $P'_A$ of size at most $25r \\alpha n$, hence $\\abs{A''} \\geq (\\tfrac 12 - 30 r \\alpha)n$. Furthermore, each vertex in $A''$ is connected to at least $(\\tfrac 12 - 40 r \\alpha)n$ vertices in $A''$. Clearly, the same holds for $B''$. \n\nBefore closing the path in both sets (hence, creating a cycle which contains all vertices that are not part of $P'_A$ or $P'_B$), which is a standard application of the blow-up lemma, we need to make sure that certain divisibility conditions hold. \nAs we wish to close the cycle by appending blocks of two layers of size $r$ (thus, $K_{{r\/2}, {r\/2}}$), we require that $\\abs{A''} \\equiv \\abs{B''} \\equiv 0 \\pmod r$. If this is the case, set $A''' = A''$ and proceed. Otherwise, if there is $0 < i < r$ such that $\\abs{A''} \\equiv i \\pmod r$, select $i$ vertices $a_1, \\ldots,a_{i} \\in A''$ and absorb them using disjoint instances $\\xi'_r(a_1), \\ldots, \\xi'_r(a_{i})$ with $D = A''$. Clearly, as this requires only finitely many vertices, such a disjoint family exists. Further, because $a_j \\in A''$, each absorber consumes $2r + 1$ vertices of $A''$, hence afterwards, the divisibility condition holds. Now, glue the absorbers sequentially to $P'_A$ using GE and sets $G_1, \\ldots,G_{i-1}$. As each gluing operation consumes $r$ vertices, the divisibility does not change hence we are left with a set $A''' = A'' \\setminus \\bc{ V(\\xi'_r(a_1)) \\cup \\ldots \\cup V(\\xi'_r(a_{i})) \\cup G_1 \\cup \\ldots \\cup G_{i-1} }$.\n\nNow it is easy to check that $(A''',B''')$ is $(\\eps,\\tfrac 12)$-super-regular and by Lemma~\\ref{lem:blowup} we find an $\\tfrac r2-$blowup of the path on all remaining vertices of $A'''$ and $B'''$ (\\ref{stepd:cover} of Section \\ref{sec:o_extremal1}). \nMoreover, the end-tuples of the path-like structure constructed before have at least $\\tfrac 12 |A'''|$ and $\\tfrac 12 |B'''|$ common neighbours in $A'''$ and $B'''$ respectively.\nTherefore, we may choose the start- and end-tuples of this path-blow-up to connect to these end-tuples.\n\nWe are left to argue that the constructed subgraph is $r$-connected and $r$-regular.\n\\begin{claim}\nThe constructed subgraph is $r$-connected and $r$-regular.\n\\end{claim}\n\\begin{claimproof}\nWhile $r$-regularity follows obviously, the $r$-connected part needs a short argument. Upon removal of up to $r-1$ bridge-vertices, the parts do not fall apart. Furthermore, removing up to $r-1$ vertices in the $\\tfrac r2$-blow-up of the path part of the subgraph does not disconnect the structure. Finally, the absorbing structure $\\xi_r$ itself is isomorphic to an $\\tfrac r2$-blowup of the path on three vertices. Moreover, disconnecting the graph by removing up to $r-1$ vertices in $\\xi_r'$ is not possible.\n\\end{claimproof}\n\n\n\n\\subsection{The odd case}\n\n\nAssume that $r$ is odd.\nThe argument in the odd case is a bit more delicate as in the even case. Indeed, while in the process above all divisibility conditions could be easily established, in the odd case, we might end with two almost cliques of odd size. If there is a set $C$, we can easily absorb those vertices in a way that after absorbing both parts of the graph contain an even number of vertices - which we require to embed a regular graph. If on the other hand there is no such set $C$, we need to be much more careful. We will tackle this problem by having two different types of bridges between $A$ and $B$, one consuming an even number of vertices of each set, one consuming an odd number - thus, depending on the size of $C$ and the parity of $A$ and $B$, we need to use two different constructions. The two types of bridges are visualised in Figure~\\ref{fig:bridge_graph_h_odd} for the special case $ r = 5$.\n\n\n\\begin{figure}\n \\centering\n \\begin{minipage}{0.45\\textwidth}\n \\begin{tikzpicture}\n \\node[circle, draw=black] (a_11) at (0, 4.5) {$a_{11}$};\n \\node[circle, draw=black] (a_12) at (0, 3) {$a_{12}$};\n \\node[circle, draw=black] (a_13) at (0, 1.5) {$a_{13}$};\n \\node[circle, draw=black] (b_11) at (0, -1.5) {$b_{11}$};\n \\node[circle, draw=black] (b_12) at (0, -3) {$b_{12}$};\n \\node[circle, draw=black] (b_13) at (0, -4.5) {$b_{13}$};\n \n \\node[circle, draw=black] (a_21) at (2, 4.5) {$a_{21}$};\n \\node[circle, draw=black] (a_22) at (2, 3) {$a_{22}$};\n \\node[circle, draw=green] (xa1) at (2, 1.5) {$x_{a_1}$};\n \\node[circle, draw=green] (xb1) at (2, -1.5) {$x_{b_1}$};\n \\node[circle, draw=black] (b_22) at (2, -3) {$b_{22}$};\n \\node[circle, draw=black] (b_23) at (2, -4.5) {$b_{23}$};\n \n \\node[circle, draw=black] (a_31) at (4, 4.5) {$a_{31}$};\n \\node[circle, draw=black] (a_32) at (4, 3) {$a_{32}$};\n \\node[circle, draw=green] (xa2) at (4, 1.5) {$x_{a_2}$};\n \\node[circle, draw=green] (xb2) at (4, -1.5) {$x_{b_2}$};\n \\node[circle, draw=black] (b_32) at (4, -3) {$b_{32}$};\n \\node[circle, draw=black] (b_33) at (4, -4.5) {$b_{33}$};\n \n \\node[circle, draw=black] (a_41) at (6, 4.5) {$a_{41}$};\n \\node[circle, draw=black] (a_42) at (6, 3) {$a_{42}$};\n \\node[circle, draw=black] (a_43) at (6, 1.5) {$a_{43}$};\n \\node[circle, draw=black] (b_41) at (6, -1.5) {$b_{41}$};\n \\node[circle, draw=black] (b_42) at (6, -3) {$b_{42}$};\n \\node[circle, draw=black] (b_43) at (6, -4.5) {$b_{43}$};\n \n \n \n \\path[-] (a_11) edge[draw=black] (a_22);\n \\path[-] (a_11) edge[draw=black] (xa1);\n \\path[-] (a_12) edge[draw=black] (xa1);\n \\path[-] (a_12) edge[draw=black] (a_21);\n \n \\path[-] (a_13) edge[draw=black] (a_21);\n \\path[-] (a_13) edge[draw=black] (a_22);\n\n \\path[-] (xa1) edge[draw=black] (a_31);\n \\path[-] (xa1) edge[draw=black] (a_32);\n \\path[-] (a_21) edge[draw=black] (xa2);\n \\path[-] (a_21) edge[draw=black] (a_31);\n \\path[-] (a_21) edge[draw=black] (a_32);\n \\path[-] (a_22) edge[draw=black] (xa2);\n \\path[-] (a_22) edge[draw=black] (a_31);\n \\path[-] (a_22) edge[draw=black] (a_32);\n \n \n \\path[-] (a_31) edge[draw=black] (a_42);\n \\path[-] (a_31) edge[draw=black] (a_43);\n \\path[-] (a_32) edge[draw=black] (a_41);\n \n \\path[-] (a_32) edge[draw=black] (a_43);\n \\path[-] (xa2) edge[draw=black] (a_41);\n \\path[-] (xa2) edge[draw=black] (a_42);\n \n \n \n \n \n \\path[-] (b_12) edge[draw=black] (b_23);\n \\path[-] (b_11) edge[draw=black] (b_22);\n \\path[-] (b_11) edge[draw=black] (b_23);\n \\path[-] (b_12) edge[draw=black] (xb1);\n \\path[-] (b_13) edge[draw=black] (b_22);\n \n \\path[-] (b_13) edge[draw=black] (xb1);\n \n \n \\path[-] (b_22) edge[draw=black] (xb2);\n \\path[-] (xb1) edge[draw=black] (b_32);\n \\path[-] (xb1) edge[draw=black] (b_33);\n \\path[-] (b_22) edge[draw=black] (b_32);\n \\path[-] (b_22) edge[draw=black] (b_33);\n \\path[-] (b_23) edge[draw=black] (b_32);\n \\path[-] (b_23) edge[draw=black] (xb2);\n \\path[-] (b_23) edge[draw=black] (b_33);\n \n \n \\path[-] (xb2) edge[draw=black] (b_43);\n \n \\path[-] (b_32) edge[draw=black] (b_43);\n \\path[-] (xb2) edge[draw=black] (b_42);\n \\path[-] (b_32) edge[draw=black] (b_41);\n \\path[-] (b_33) edge[draw=black] (b_41);\n \\path[-] (b_33) edge[draw=black] (b_42);\n \n \n \\path[-] (xb2) edge[draw=green] (xa2);\n \\path[-] (xb1) edge[draw=green] (xa1);\n\n \\end{tikzpicture}\n \\end{minipage} \\hfill\\vline\\hfill\n \\begin{minipage}{0.45\\textwidth}\n \\begin{tikzpicture}\n \\node[circle, draw=black] (a_11) at (0, 4.5) {$a_{11}$};\n \\node[circle, draw=black] (a_12) at (0, 3) {$a_{12}$};\n \\node[circle, draw=black] (a_13) at (0, 1.5) {$a_{13}$};\n \\node[circle, draw=black] (b_11) at (0, -1.5) {$b_{11}$};\n \\node[circle, draw=black] (b_12) at (0, -3) {$b_{12}$};\n \\node[circle, draw=black] (b_13) at (0, -4.5) {$b_{13}$};\n \n \\node[circle, draw=black] (a_21) at (2, 4.5) {$a_{21}$};\n \\node[circle, draw=black] (a_22) at (2, 3) {$a_{22}$};\n \\node[circle, draw=black] (a_23) at (2, 1.5) {$a_{23}$};\n \\node[circle, draw=black] (b_21) at (2, -1.5) {$b_{21}$};\n \\node[circle, draw=black] (b_22) at (2, -3) {$b_{22}$};\n \\node[circle, draw=black] (b_23) at (2, -4.5) {$b_{23}$};\n \n \\node[circle, draw=black] (a_31) at (4, 4.5) {$a_{31}$};\n \\node[circle, draw=black] (a_32) at (4, 3) {$a_{32}$};\n \\node[circle, draw=black] (a_33) at (4, 1.5) {$a_{33}$};\n \\node[circle, draw=black] (b_31) at (4, -1.5) {$b_{31}$};\n \\node[circle, draw=black] (b_32) at (4, -3) {$b_{32}$};\n \\node[circle, draw=black] (b_33) at (4, -4.5) {$b_{33}$};\n \n \\node[circle, draw=black] (a_41) at (6, 4.5) {$a_{41}$};\n \\node[circle, draw=black] (a_42) at (6, 3) {$a_{42}$};\n \\node[circle, draw=black] (a_43) at (6, 1.5) {$a_{43}$};\n \\node[circle, draw=black] (b_41) at (6, -1.5) {$b_{41}$};\n \\node[circle, draw=black] (b_42) at (6, -3) {$b_{42}$};\n \\node[circle, draw=black] (b_43) at (6, -4.5) {$b_{43}$};\n \n \\node[circle, draw=green] (xa) at (3, 2.25) {$x_{a}$};\n \\node[circle, draw=green] (xb) at (3, -2.25) {$x_{b}$};\n \n \n \n \\path[-] (a_12) edge[draw=black] (a_23);\n \\path[-] (a_11) edge[draw=black] (a_23);\n \n \\path[-] (a_11) edge[draw=black] (a_22);\n \n \n \\path[-] (a_12) edge[draw=black] (a_21);\n \n \\path[-] (a_13) edge[draw=black] (a_21);\n \\path[-] (a_13) edge[draw=black] (a_22);\n\n\n \\path[-] (a_21) edge[draw=black] (a_31);\n \\path[-] (a_21) edge[draw=black] (a_32);\n \\path[-] (a_22) edge[draw=black] (a_31);\n \\path[-] (a_22) edge[draw=black] (a_32);\n \n \n \\path[-] (a_31) edge[draw=black] (a_42);\n \\path[-] (a_31) edge[draw=black] (a_43);\n \\path[-] (a_32) edge[draw=black] (a_41);\n \n \\path[-] (a_32) edge[draw=black] (a_43);\n \\path[-] (xa2) edge[draw=black] (a_41);\n \\path[-] (xa2) edge[draw=black] (a_42);\n \n \n \n \n \n \\path[-] (b_12) edge[draw=black] (b_23);\n \\path[-] (b_11) edge[draw=black] (b_22);\n \\path[-] (b_11) edge[draw=black] (b_23);\n \\path[-] (b_12) edge[draw=black] (xb1);\n \\path[-] (b_13) edge[draw=black] (b_22);\n \n \\path[-] (b_13) edge[draw=black] (xb1);\n \n \n \n \n \n \\path[-] (b_22) edge[draw=black] (b_32);\n \\path[-] (b_22) edge[draw=black] (b_33);\n \\path[-] (b_23) edge[draw=black] (b_32);\n \n \\path[-] (b_23) edge[draw=black] (b_33);\n \n \n \n \n \\path[-] (b_32) edge[draw=black] (b_43);\n \n \\path[-] (b_32) edge[draw=black] (b_41);\n \\path[-] (b_33) edge[draw=black] (b_41);\n \\path[-] (b_33) edge[draw=black] (b_42);\n \n \n \\path[-] (b_31) edge[draw=black] (b_42);\n \\path[-] (b_31) edge[draw=black] (b_43);\n \n \\path[-] (b_21) edge[draw=black, bend right=10] (b_33);\n \\path[-] (b_31) edge[draw=black, bend left=10] (b_23);\n \\path[-] (b_31) edge[draw=black] (b_21);\n \n \\path[-] (xb) edge[draw=black] (b_21);\n \\path[-] (xb) edge[draw=black] (b_22);\n \\path[-] (xb) edge[draw=black] (b_31);\n \\path[-] (xb) edge[draw=black] (b_32);\n \n \\path[-] (a_23) edge[draw=black, bend left=10] (a_31);\n \\path[-] (a_33) edge[draw=black, bend right=10] (a_21);\n \\path[-] (a_23) edge[draw=black] (a_33);\n \n \\path[-] (xa) edge[draw=black] (a_22);\n \\path[-] (xa) edge[draw=black] (a_32);\n \\path[-] (xa) edge[draw=black] (a_23);\n \\path[-] (xa) edge[draw=black] (a_33);\n \n \n \\path[-] (xa) edge[draw=green] (xb);\n \n\n \\end{tikzpicture}\n \\end{minipage}\n \\caption{The two types of connections between the sets $A$ and $B$ in the special case $r = 5$.}\n \\label{fig:bridge_graph_h_odd}\n\\end{figure}\n\nWe begin by showing that we find three pairs of bridges of the first type, using an even number of vertices of both classes (\\ref{stepd:bridge} of Section \\ref{sec:o_extremal1}).\n\n\\begin{claim}\\label{lemma_matching_edges_odd}\nSuppose $\\delta(G) \\geq \\tfrac{n+r-2}{2}$ and $\\abs{A} \\leq \\abs{B}$. Furthermore, let $n$ be large enough. There is a matching $(x_{a_1}x_{b_1}, \\ldots, x_{a_{r+1}}x_{b_{r+1}})$ such that $\\abs{N(x_{a_i}) \\cap A} \\geq \\tfrac n5$ and $\\abs{N(x_{b_j}) \\cap B} \\geq \\tfrac n5$ for all $i, j \\leq r+1$.\n\\end{claim}\n\\begin{claimproof}\nAs in the proof of Claim~\\ref{lemma_matching_edges}, it suffices to find $r+1$ edges from $A$ to $V \\setminus A$. \n\nFirst, suppose that $n$ is odd. If $\\abs{A} \\leq \\tfrac{n-r-2}{2}$, the minimum degree of $ \\tfrac{n+r-2}{2}$ guarantees that each vertex of $A$ needs to find at least $\\tfrac{n+r-2}{2} - (\\tfrac{n-r-2}{2} - 1) = r+1$ neighbors outside of $A$, hence the assertion follows.\nSuppose $\\abs{A} = \\tfrac{n-r}{2} + i$ with $i = 1, \\ldots,\\floor{\\tfrac r2}$. In this case, $\\abs{V \\setminus A} = \\tfrac{n+r}{2} - i$. Each vertex of $A$ finds at least $r-i$ neighbors outside of $A$. Suppose that $N(A) \\setminus A$ has size at most $r$ (thus, all edges from $A$ into the rest of the graph belong to at most $r$ vertices). Now pick a different vertex in the complement (which exists, as $\\tfrac{n+r}{2} - i \\gg r$). This vertex requires $\\tfrac{n+r-2}{2} - (\\tfrac{n+r}{2} - i - 1) = i$ neighbors in $A$, which is a contradiction.\nIf finally $\\abs{A} = \\tfrac{n-r}{2}$, each vertex of $A$ has at least $r$ neighbors in the complement of $A$. If all vertices of $A$ share those $r$ vertices (hence, we only find $r$ matching edges), those vertices are connected to all vertices in $A$, hence can be moved to $A$ by only increasing $\\alpha$-extremity, thus the assertion follows from the previous case.\n\nNow, if $n$ is even, because the minimum degree needs to be an integer, it is at least $\\tfrac{n+1+r-2}{2}$, hence upon removal of one vertex, we are left with a graph on $n' = n-1$ vertices and minimum degree at least $\\tfrac{n+1+r-2}{2} -1= \\tfrac{n'+r-2}{2}$ (and $n'$ being odd). Hence the assertion follows from the previous discussion.\n\\end{claimproof}\n\nSimilarly as in the even case, Claim~\\ref{lemma_matching_edges_odd} gives us $\\tfrac{r+1}{2}$ pairs of bridge-edges as in Figure \\ref{fig:bridge_graph_h_odd} (the green part). Clearly, the rest of the bridge graph can be created completely analogously to Claim~\\ref{bridge_graph_existence}.\n\nBy the same token we get the following claim immediately.\n\\begin{claim}\\label{cor_matching_edges_odd}\nSuppose $\\delta(G) \\geq \\tfrac{n+r-2}{2}$ and $\\abs{A} \\leq \\abs{B}$. Furthermore, let $n$ be large enough. There is a matching $(x_{a_1}x_{b_1}, \\ldots, x_{a_{r-1}}x_{b_{r-1}})$ such that $\\abs{N(x_{a_i}) \\cap A} \\geq \\tfrac n5$ and $\\abs{N(x_{b_j}) \\cap B} \\geq \\tfrac n5$ for all $i, j \\leq r-1$. Furthermore, there are different vertices $x_a \\in A, x_b \\in B$ such that $x_ax_b \\in E(G)$, $\\abs{N(x_a) \\cap A} \\geq r-1$ and $\\abs{ N(x_b) \\cap B} \\geq r-1$.\n\\end{claim}\n\\begin{claimproof}\nThis follows directly from the proof of Claim~\\ref{lemma_matching_edges_odd}.\n\\end{claimproof}\n\nNext, we re-define the gluing operation GE to GO as follows.\n\\begin{claim}[Gluing operation GO]\nGiven two disjoint sets $D_1, D_2 \\subset A'_0$ of sizes exactly $\\tfrac{r+1}{2}$, we find two disjoint sets $D'_1, D'_2 \\subset A'_0 \\setminus (D_1 \\cup D_2)$ of size $\\tfrac{r+1}{2}$ such that\n\\[G[D_1, D'_1] \\equiv G[D'_1, D'_2] \\equiv G[D'_2, D_2] \\equiv K_{(r+1)\/2, (r+1)\/2} \\, .\\]\n\\end{claim}\n\\begin{claimproof}\nThis follows directly from the fact that each vertex in $A'_0$ is connected to at least $(\\tfrac 12 - 4 \\alpha)n$ vertices in $A'_0$ and GO uses only finitely many vertices of the neighborhoods.\n\\end{claimproof}\nWe stress at this point that GO can be applied to blocks whose end-vertices $D_1$ have currently degree $\\tfrac{r-1}{2}$ (then we chose all edges from the connecting graphs $K_{(r+1)\/2, (r+1)\/2}$) or degree $\\tfrac{r+1}{2}$ (then we chose the complete bipartite graph between $D'_1$ and $D'_2$ and at the other connections, we remove one matching of size $\\tfrac{r+1}{2}$.\nFurthermore observe, that gluing consumes $r+1$ vertices from the underlying set.\n\nWe proceed as follows. If $\\abs{C} > 0$ or $\\abs{C} = 0$ and $\\abs{A'_0}, \\abs{B'_0} \\equiv 0 \\pmod 2$, we create $\\tfrac{r+1}{2}$ pairs of bridge vertices by Claim~\\ref{lemma_matching_edges_odd}. Otherwise, we create $\\tfrac{r-1}{2}$ pairs of bridge vertices and one additional bridge by Claim~\\ref{cor_matching_edges_odd}. In both cases, we glue the bridges in $A$ and $B$ respectively together using mutually disjoint vertex sets $D^A_1, \\ldots,D^A_{(r-1)\/2}$ and $D^B_1, \\ldots,D^B_{(r-1)\/2}$ constructing $P_A, P_B$ and, similarly as in the even case, we set $A' = A'_0 \\setminus V(P_A), B' = B'_0 \\setminus V(P_B)$. \nClearly, the parity of $A'$ and $B'$ are both even.\n\nNext, we define absorbing structures for the left-over vertices and for absorbing vertices in order to guarantee divisibility (\\ref{stepd:absorb} of Section \\ref{sec:o_extremal1}). They need to be defined slightly differently as in the even case (Figure \\ref{fig:absorber_odd}).\n\n\\begin{definition}\n Let $D \\in \\cbc{A', B'}$ and $u$ a vertex such that $\\abs{ N(u) \\cap D } \\geq \\tfrac n6$. Define $\\xi_r(u)$ as follows.\n \\begin{itemize}\n \\item Select $D_1 = \\cbc{d_1, d_2, \\ldots, d_{(r+1)\/2}}, D_2 = \\cbc{d'_1, \\ldots, d'_{(r+1)\/2}} \\subset N(u) \\cap D$, hence $r$ pairwise disjoint vertices.\n \\item Select $D' = \\cbc{ u'_1, \\ldots, u'_{(r-1)\/2}} \\subset N(D_1) \\cap N(D_2) \\cap D \\setminus (D_1 \\cup D_2 \\cup \\cbc{u}) $.\n \\item Select $E_0 = \\cbc{ e_1, \\ldots, e_{(r+1)\/2} }$ in the joint $D$-neighborhood of $D_1 \\setminus (D_2 \\cup D' \\cup \\cbc{u})$. \n \\end{itemize}\n Define $\\xi_r(u)$ as the graph containing $E_0, D_1, D_2, D'$ and $u$ as well as all the edges from $D_1$ to $D' \\cup \\cbc{u}$. Furthermore, take all edges from $D_2$ to $D' \\cup \\cbc{u}$ removing one matching of size $\\tfrac{r+1}{2}$ and the edges from $E_0$ to $D_1$ removing a matching as well.\n Furthermore, for two adjacent vertices $u_1, u_2 \\in D$ define $\\xi'_r(u_1, u_2)$ via\n \\begin{itemize}\n \\item Select $F_1 = \\cbc{f_1, f_2, \\ldots, f_{(r+1)\/2}}, F' = \\cbc{f'_1, \\ldots, f'_{(r+1)\/2}} \\subset N(u_1) \\cap N(u_2) \\cap D$, hence $r+1$ pairwise disjoint vertices in the joint neighborhood of $u_1$ and $u_2$.\n \\item Select the following edges\n \\begin{itemize}\n \\item $f_1u_1, \\ldots, f_{(r-1)\/2}u_1$,\n \\item $f'_1u_1, \\ldots, f'_{(r-1)\/2}u_1$,\n \\item $f_2u_2, \\ldots, f_{(r+1)\/2}u_2$,\n \\item $f'_2u_2, \\ldots, f'_{(r+1)\/2}u_2$ and\n \\item $f_1f'_1$ as well as $f_{(r+1)\/2}f'_{(r+1)\/2}$.\n \\end{itemize}\n \\item Finally, draw $(r-2)$ half-edges at each $f_i, f'_i$ and match them such that a simple graph is induced by the matching.\n \\end{itemize}\n\\end{definition}\nAs $u_1, u_2 \\in D$, hence the neighborhood of $u_1$, $u_2$ contains only vertices that, themselves, have $(\\tfrac 12 - 30 r \\alpha)n$ vertices from $D$, the joint neighborhood of those two vertices has size at least $(\\tfrac 12 - 60 r \\alpha )n$, $\\xi'(u_1, u_2)$ is well defined. Observe that absorbing a vertex $u \\not \\in D$ consumes $2r + 1$ vertices from $D$ while absorbing $u_1, u_2 \\in D$ consumes $r+3$ vertices (including $u_1, u_2$) in $D$.\n\nAs in the even case, we find a family of disjoint structures to absorb at least $100 \\alpha n$ different vertices.\n\n\\begin{figure}\n \\centering\n \\begin{tikzpicture}[scale=0.8]\n \n \\node[circle, draw=blue] (e1) at (0, 3) {$e_1$};\n \\node[circle, draw=blue] (e2) at (0, 1) {$e_2$};\n \\node[circle, draw=blue] (e3) at (0, -1) {$e_3$};\n\n \\node[circle, draw=green!40] (d1) at (2, 3) {$d_1$};\n \\node[circle, draw=green!40] (d2) at (2, 1) {$d_2$};\n \\node[circle, draw=green!40] (d3) at (2, -1) {$d_3$};\n\n \\node[circle, draw=green!40] (dp1) at (6, 3) {$d'_{1}$};\n \\node[circle, draw=green!40] (dp2) at (6, 1) {$d'_{2}$};\n \\node[circle, draw=green!40] (dp3) at (6, -1) {$d'_{3}$};\n \n \\node[circle, draw=blue!40] (up1) at (4, 3) {$u'_{1}$};\n \\node[circle, draw=blue!40] (up2) at (4, 1) {$u'_{2}$};\n \\node[circle, draw=green] (u) at (4, -1) {$u$};\n\n \n \\path[-] (u) edge[draw=green] (d1);\n \\path[-] (u) edge[draw=green] (d2);\n \\path[-] (u) edge[draw=green] (d3);\n \\path[-] (u) edge[draw=green] (dp1);\n \\path[-] (u) edge[draw=green] (dp2);\n \n \n \\path[-] (d1) edge[draw=blue] (up1);\n \\path[-] (d2) edge[draw=blue] (up1);\n \\path[-] (d1) edge[draw=blue] (up2);\n \\path[-] (d2) edge[draw=blue] (up2);\n \\path[-] (d3) edge[draw=blue] (up1);\n \\path[-] (d3) edge[draw=blue] (up2);\n \n \n \\path[-] (d2) edge[draw=blue] (e1);\n \\path[-] (d3) edge[draw=blue] (e1);\n \\path[-] (d1) edge[draw=blue] (e2);\n \n \\path[-] (d3) edge[draw=blue] (e2);\n \\path[-] (d1) edge[draw=blue] (e3);\n \\path[-] (d2) edge[draw=blue] (e3);\n \n \n \n \n \\path[-] (dp2) edge[draw=blue] (up1);\n \\path[-] (dp3) edge[draw=blue] (up1);\n \\path[-] (dp1) edge[draw=blue] (up2);\n \n \\path[-] (dp3) edge[draw=blue] (up2);\n \n \n \n \\node (h1) at (7, 4) {};\n \\node (h2) at (7, -2) {};\n \n \n \\node[circle, draw=green!40] (f1) at (8, 3) {$f_1$};\n \\node[circle, draw=green!40] (f2) at (8, 1) {$f_2$};\n \\node[circle, draw=green!40] (f3) at (8, -1) {$f_3$};\n \n \\node[circle, draw=green!40] (u1) at (10, 2) {$u_1$};\n \\node[circle, draw=green!40] (u2) at (10, 0) {$u_2$};\n \n \\node[circle, draw=green!40] (fp1) at (12, 3) {$f'_1$};\n \\node[circle, draw=green!40] (fp2) at (12, 1) {$f'_2$};\n \\node[circle, draw=green!40] (fp3) at (12, -1) {$f'_3$};\n\n \\path[-] (h1) edge[draw=black] (h2);\n \n \\path[-] (fp1) edge[draw=green] (u1);\n \\path[-] (fp2) edge[draw=green] (u2);\n \\path[-] (u2) edge[draw=green] (u1);\n \n \\path[-] (fp2) edge[draw=green] (u1);\n \\path[-] (fp3) edge[draw=green] (u2);\n \\path[-] (f1) edge[draw=green] (u1);\n \\path[-] (f2) edge[draw=green] (u2);\n \n \\path[-] (f2) edge[draw=green] (u1);\n \\path[-] (f3) edge[draw=green] (u2);\n \n \\path[-] (fp1) edge[draw=green] (f1);\n \n \n \n \n \n \n \n \\path[-] (fp3) edge[draw=green] (f3);\n \n \n \\end{tikzpicture}\n \n \\caption{Absorbers $\\xi_5(u)$ (left) and $\\xi'_5(u_1, u_2)$ (right) with $r=5$, where $u, u_1, u_2$ are the green vertices, the green vertices are inside the $A'$-neighborhood of $u$ (or $u_1, u_2$ respectively) and the blue vertices are vertices chosen from $A'$.}\n \\label{fig:absorber_odd}\n\\end{figure}\n\n\nSubsequently, we absorb $C$ using independent copies of $\\xi_r(\\cdot)$ such that the parity of the remaining vertices in the almost cliques is even. Then, as above, we glue the absorbers together by GO and are left with sets $A''$ and $B''$ (hence, $A'$ without the glued structures) and the path-like subgraph $P'_A$ of size at most $25r \\alpha n$, hence $\\abs{A''} \\geq (\\tfrac 12 - 30 r \\alpha n)$. As each vertex of $C$ has degree at least $\\tfrac{\\alpha}{100} n$ into $A''$ and $B''$, we can absorb one vertex such that in the end $A''$ and $B''$ contain an even number of vertices. Furthermore, each vertex in $A''$ is connected to at least $(\\tfrac 12 - 40 r \\alpha)n$ vertices in $A''$ and the same applies to $B''$. \n\nAgain, as in the even case, we need to make sure that $\\abs{A''} \\equiv 0 \\pmod{2r}$, as we want to close the cycle by blocks of subsequently fully connected layers of sizes $\\tfrac{r+1}{2}$ where the second block misses one matching of size $r+1$. If the divisibility condition holds, set $A''' = A''$ and proceed. Otherwise, if there is $0 < i < 2r$ such that $\\abs{A''} \\equiv i \\pmod{2r}$, $i$ has to be even as the parity of $A''$ guarantees. Select $\\tfrac i2$ pairs vertices $(a_{11}, a_{12}), \\ldots,(a_{i\/2, 1} a_{i\/2, 2}) \\in A'' \\times A''$ and absorb them using disjoint instances $\\xi'_r(a_{11}, a_{12}), \\ldots, \\xi'_r(a_{i\/2, 1} a_{i\/2, 2})$ with $D = A''$. As each absorber consumes $r + 3$ vertices, the divisibility condition now holds. Finally, as in the even case, glue the absorbed parts together with GO which does not change the divisibility by $r+1$. Thus, we are left with a set $A'''$ which consists of the vertices of $A''$ without the absorbed vertices and the gluing structures. Analogously, the same applies for $B'''$. Now, as in the even case, the result directly follows from Lemma~\\ref{lem:blowup} and the following claim (\\ref{stepd:cover} of Section \\ref{sec:o_extremal1}).\n\n\\begin{claim}\nThe constructed subgraph is $r$-connected and $r$-regular.\n\\end{claim}\n\\begin{claimproof}\nAs in the even case, $r$-regularity as well as $r$-connectivity on the $\\tfrac{r+1}{2}$-blow-up of the path part is obvious.\nThe first type of bridge (build with Claim~\\ref{lemma_matching_edges_odd}, see Figure~\\ref{fig:bridge_graph_h_odd} on the left) does not harm connectivity as before. In the second type of bridge (build with Claim~\\ref{cor_matching_edges_odd}, see Figure~\\ref{fig:bridge_graph_h_odd} on the right) only the special vertices $x_a$ and $x_b$ need our attention. But as they are of degree $r$ and connected to $(\\tfrac{r-1}{2}$ vertices on both sides of the $K_{(r+1)\/2), (r+1)\/2}$, isolating a part of the graph is not possible either. The absorbing structures clearly sustain the connectivity property.\n\\end{claimproof}\nThis finishes the proof of the first extremal case.\n\\end{proof}\n\n\\@ifstar{\\origsection*}{\\mysection}{Extremal Case II}\n\\label{sec:extremal2}\nIn this section we deal with the second extremal case and follow~\\ref{stepe:find}--\\ref{stepe:cover} as outlined in Section~\\ref{sec:o_extremal2}.\nWe start by proving the auxiliary lemma for finding stars.\n\n\\begin{proof}[Proof of Lemma~\\ref{lem:stars}]\nLet $\\alpha = \\tfrac{1}{32s^2(s+1)}$.\nAssume we have already found $0 \\le t < 2m$ copies of $K_{1,s}$ and let $V'$ be the remaining vertices.\nThen, by the maximum degree condition in $G$,\n\\[e(G[V']) \\ge \\tfrac 12 n (m+s-1) - t (s+1) 4 s \\alpha n \\,.\\]\nIf $m \\ge s+1$ this is at least $\\tfrac 14 n (m+s-1) \\ge \\tfrac 12 ns$ and gives a vertex of degree at least $s$ in $G[V']$.\nOn the other hand, if $m \\le s$ the above is at least $(\\tfrac 12 s - \\tfrac 14) n > \\tfrac{s-1}{2}n$ and again this gives a vertex of degree at least $s$ in $G[V']$.\n\\end{proof}\n\n\\begin{proof}[Proof of Extremal Case II]\nLet $r \\ge 2$ and $s = \\lceil \\tfrac{r}{2} \\rceil \\ge 1$.\nLet $\\eps>0$ be given by Lemma~\\ref{lem:blowup} on input $\\tfrac 12$, $\\tfrac 12$, and $r$.\nWe obtain $\\alpha>0$ from Lemma~\\ref{lem:stars} and additionally assume that $40 s^2 \\alpha < \\eps$.\nThen let $G$ be an $n$-vertex graph with minimum degree $\\delta(G) \\ge \\tfrac{n+r-2}{2}$ and $nr \\equiv 0 \\pmod 2$.\nFurther assume, that there is a partition of $V(G)$ into $A$ and $B$ of size $|A|+m=\\tfrac 12 n=|B|-m$, where $0 \\le m \\le \\alpha n$ such that between these sets we have minimum degree $\\alpha n$ and all but at most $\\alpha n$ vertices from $A$ (or $B$) have degree $|B|-\\alpha n$ (or $|A|-\\alpha n$) into $B$ (or $A$).\n\n\\noindent \\textbf{\\ref{stepe:find}.}\nNote $\\delta(G[B]) \\ge m+s-1$.\nLet $B' \\subseteq B$ be the vertices of degree at most $2 s \\alpha n$ in $G[B]$ and let $m'=|B \\setminus B'|$.\nIf $m' \\norm{r}$. Consequently, there are no\ninfinite reduction sequences, or, in other words, the term rewriting\nsystem is strongly terminating.\nNotice that because we consider term rewriting modulo commutativity of\n$\\OR$ and $\\AND$, we have to verify that the left-hand side and the right-hand\nside of equations \\eqref{E1} and \\eqref{E2}\nhave equal norms~\\cite{KBV2001}. This is clearly the\ncase.\n\\item \nNow we prove that the terms \nproduced by the grammar~\\eqref{grammar:NF} \nare exactly the normal forms with respect to $\\thetrs$.\nFor the terms in $\\asterms$, none of the rewrite rules can be applied,\nbecause these terms do not contain $\\OR$, have no occurrences of\n$\\AND$ containing an argument of type $\\AND$, have no occurrences of\n$\\SAND$ containing an argument of type $\\SAND$, and do not contain\noperators with a single argument.\nWe extend this to terms $\\nfterms$ by observing that all $\\OR$\noperators occurring in such terms have at least two arguments and that\nall these arguments are different.\n\nConversely, consider a term $t$ in normal form that \ncontains an $\\OR$ operator. Then $t=\\OR(t_1,\\ldots,t_n)$, where the $t_i$\ndo not contain an $\\OR$ operator, else \\eqref{E3}, \\eqref{E10}, or\n\\eqref{E10'} can be applied. \nIt remains to show that normal form terms without occurrence of an\n$\\OR$ operator are in $\\asterms$. Such terms are basic terms or have\n$\\SAND$ or $\\AND$ as their top-level operator. The last two cases are\nsymmetric and we therefore only consider the case\n$\\AND(t_1,\\ldots,t_n)$. We must show that each $t_i$ is a basic term\nor in the form $t_i=\\SAND(t_1',\\ldots,t_m')$. Suppose not, then there\nexists a $t_i$ that has $\\AND$ as its top-level operator. It follows\nthat the term is not in normal form because \\eqref{E4} can be applied.\n\n\\item The normal forms are unique. \nTo show that the normal forms are unique, assume that \n$\\nfterm_1$ and $\\nfterm_2$ are both normal forms for a\n \\SAND~attack tree $t$. \nSince the rewrite system $\\thetrs$ \nwas constructed by orienting the axioms from $\\ESP$, \nwe have that $\\ESP\\vdash \\nfterm_1=\\nfterm_2$. \nThis means that $\\sem{\\nfterm_1}=\\sem{\\nfterm_2}$.\nFrom bijectivity proven in Lemma~\\ref{lem:bijection}, we obtain \n$\\nfterm_1 = \\nfterm_2$.\n\n\\item Now that we have proven termination and uniqueness of normal forms, it\nimmediately follows that the term rewriting system is\nconfluent~\\cite{D2005}.\n\\end{enumerate} \n\\qed\n\\end{proof}\n \n\n\n\n\nExample~\\ref{ex:nf} illustrates the notion of \ncanonical form for \\SAND~attack trees.\n}\n\n\\begin{example}\n\\label{ex:nf}\nThe canonical form of the \\SAND~attack tree $t$ \nin Figure~\\ref{fig:attack-tree} is the tree \n\\[\nt' = \\OR\\Big(\\SAND\\big(\\ftprhostsx, \\rshx, \\localbofx\\big), \\\\\n \\AND\\big(\\sshbofx, \\rsarefbofx\\big)\\Big)\n\\]\nshown in Figure~\\ref{fig:attack-tree2}. \nIt \nis easily seen to be in normal form with respect to $\\thetrs$.\n\\end{example}\n\\begin{figure}[h]\n\\centering\n\\includegraphics[height=3.4cm]{atree3}\n\\caption{\\SAND\\ attack tree $t'$ equivalent to \\SAND\\ attack tree $t$ from \nFigure~\\ref{fig:attack-tree}}\n\\label{fig:attack-tree2}\n\\end{figure}\n\n\\subsection{SP Semantics as a Generalization of the Multiset Semantics}\n\\label{sec:SP_vs_M}\n\nHaving a complete set of axioms for the SP semantics\nallows us to formalize the relation between \n\\SAND\\ attack trees under the SP semantics and \nattack trees under the multiset semantics, denoted by $\\msem{\\cdot}$. \nThis is achieved by extracting \na complete set of axioms for \nthe multiset semantics for \nattack trees from \nthe set $\\ESP$. \nLet $\\ESM$ be the subset of axioms from $\\ESP$ that\ndo not contain the $\\SAND$ operator, \\ie, \n$\\ESM=\\{$\\eqref{E1}, \\eqref{E2}, \\eqref{E3}, \\eqref{E4}, \\eqref{E5},\n\\eqref{E6}, \\eqref{E10}, \\eqref{E11}$\\}$.\n\\begin{theorem}\n\\label{th:multiset-axioms}\nThe axiom system $\\ESM$\nis a complete set of axioms for the multiset\nsemantics for attack trees.\n\\end{theorem}\n{\n\\begin{proof}\nIn \\cite[Theorem 4.9]{KoMaRaSc_JLC}, a complete axiomatization of\nthe multiset semantics for an extention of attack trees \ncalled attack--defense trees (ADTrees) is given. In the following, we\ncall that axiomatization $\\EADT$.\nADTrees are a superset of attack trees.\nThey may contain defender's nodes \nmodeled by the so called\nopponent's functions and countermeasures. \nWe claim that $\\ESM$\nis a complete axiomatization of the multiset semantics for attack trees.\nObviously if two attack trees\nare equal with respect to $\\ESM$, then they\nare also equal with respect to $\\EADT$.\nThis is clear, because $\\ESM\\subset \\EADT$.\n\nConversely, we prove that if two attack trees are equal with respect to \n$\\EADT$, they are equal with respect to $\\ESM$. This\nfollows from the following syntactical reasoning. \n$\\EADT$ contains function symbols which we call\n\\emph{countermeasures}. \nObserve by inspecting the axioms of $\\EADT$ that if a\ncountermeasure occurs at the left-hand side of an equation, then it also\noccurs at the right-hand side, and vice versa. Therefore, axioms \n$(E_{13}),\n(E_{16}), (E_{17}), (E_{18}), (E_{19}), (E_{20})$ from $\\EADT$\ncan never be used in a derivation of equality of\ntwo standard attack trees. Further, observe that the remaining \naxioms $(E_9)$ and $(E_{12})$ from $\\EADT$ make use of \n\\emph{opponent's functions}.\nIn these axioms, \nan opponent function occurs on the left-hand side if and only\nif it occurs on the right hand side. Thus \nthese axioms are never used to equate\ntwo attack trees which do not contain opponent's nodes. \nThe remaining axioms are precisely \n\\eqref{E1}, \\eqref{E2}, \\eqref{E3}, \\eqref{E4}, \\eqref{E5},\n\\eqref{E6}, \\eqref{E10}, \\eqref{E11}.\nSo, we can only use these axioms \nto derive equalities\nof attack trees with respect to $\\EADT$, which implies that such a\nderivation is also possible using axioms from $\\ESM$.\n\\qed\n\\end{proof}\n}\n\n\nBy comparing the complete sets of axioms \n$\\ESP$ and $\\ESM$ we obtain that two attack trees are equivalent under\nthe multiset semantics if and only if they are equivalent under the SP \nsemantics. This is formalized in the following theorem. \n\\begin{theorem}\n\\label{th:SP_vs_M}\n$\\SAND$ attack trees under the SP semantics are a \\emph{conservative\nextension} of attack trees under the multiset semantics.\n\\end{theorem}\n\\begin{proof}\n{\nLet $t$ and $t'$ be standard attack trees.\nLet $\\msem{t}$ and $\\msem{t'}$ be their interpretation in the \nmultiset semantics and $\\sem{t}$ and\n$\\sem{t'}$ be their interpretation in the SP semantics.\nWe prove that $\\msem{t} = \\msem{t'}$ if and only if $\\sem{t} =\n\\sem{t'}$.\n\nBy Theorem~\\ref{th:multiset-axioms}, a complete axiomatization of \nthe multiset semantics for attack \ntrees \nconsists of axioms~\\eqref{E1}, \\eqref{E2}, \\eqref{E3}, \\eqref\n{E4}, \\eqref{E5}, \\eqref{E6}, \\eqref{E10}, \\eqref{E11}.\nThe complete axiomatization of the SP semantic for \n$\\SAND$ attack trees additionally \ncontains axioms \\eqref{E4'}, \\eqref{E6'},\nand \\eqref{E10'}. Thus, every equivalence of attack\ntrees under the multiset semantics \nis clearly an equivalence of $\\SAND$ attack trees\nunder the SP semantics.\n\nTo see the converse, \nwe show that the additional axioms do not introduce new\nequalities on standard attack trees. \nFirst inspect the three additional axioms and note that all of them\ncontain the $\\SAND$ operator.\n\nNext, observe that for all axioms, the set of variables occurring on the\nleft-hand side is equal to the set of variables occurring on the\nright-hand side. Thus, there is no axiom eliminating all\noccurrences of a variable. \nIn particular, we claim that \nall axioms transform \nterms containing a $p$-ary $\\SAND$ expression, where $p\\geq 2$, into\nterms containing a $q$-ary $\\SAND$ expression, for some $q\\geq 2$. \nThis is evident for equations without the $\\SAND$ operator (since no\nvariables are eliminated) and remains\nto be shown for equations \\eqref{E4'}, \\eqref{E6'},\nand \\eqref{E10'}. \nAxiom~\\eqref{E6'} introduces and removes unary $\\SAND$, but does not\nmodify the single variable $A$ and therefore satisfies the claim. \nThe arities of the two left-hand side $\\SAND$ operators \nin equation~\\eqref{E4'} are $l$ and $k+l+m$ and the arity of the \nright-hand side operator is $k+l+m$, where $k,m\\geq 0$ and $l\\geq 1$.\nSince $1\\leq l\\leq k+l+m$ and both sides contain a $\\SAND$ operator of\narity $k+l+m$, if either of the two sides\ncontains a $\\SAND$ operator with two or more arguments, then so does\nthe other side.\nFinally, since $l\\geq 1$, \nthe arity of the $\\SAND$ operator on the left-hand side of\nequation~\\eqref{E10'} is equal to the arities of the $\\SAND$ operators\non its right-hand side and at least one $\\SAND$ operator\noccurs on the right-hand side.\n\nWe can now show that none of the three axioms~\\eqref{E4'}, \\eqref{E6'}, \\eqref\n{E10'} introduces new equalities on\nstandard attack trees. \nIn particular, axiom~\\eqref{E6'} introduces and removes unary\n$\\SAND$, but this does not introduce new equalities on standard attack\ntrees. Equations~\\eqref{E4'} and~\\eqref{E10'} match unary $\\SAND$, but\nrequire a further $\\SAND$ with $2$ or more arguments to add a new equality.\nSince, by the above claim, no $p$-ary $\\SAND$ for $p\\geq 2$ \ncan be introduced with any of\nthe equations, the additional equations do not introduce new\nequalities on standard attack trees.\n\\qed\n}\n\n\\end{proof}\n\n{\n}\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Attributes}\n\\label{sec:attributes}\n\n\nAttack trees do not only serve to represent security scenarios \nin a graphical way. They can also be used to \nquantify such scenarios with respect to \na given parameter, called an \\emph{attribute}. \nTypical examples of attributes include\nthe likelihood that the attacker's goal is satisfied and \nthe minimal time or cost of an attack. \nSchneier described~\\cite{Schn} \nan intuitive bottom-up algorithm for calculating attribute values\non attack trees:\nattribute values are assigned to the leaf nodes \nand two functions\\footnote{These are actually \nfamilies of functions representing infinitely many $k$-ary function symbols,\nfor all $k \\geq 2$.} (one for the \\OR\\ and one for the \\AND\\ refinement)\nare used to propagate the attribute value up to the root node. \nMauw and Oostdijk showed~\\cite{MaOo} that \nif the binary operations induced by the two functions \ndefine a semiring, \nthen the \nevaluation of the attribute on two attack trees \nequivalent with respect to the multiset semantics \nyields the same value. This result has \nbeen generalized to any semantics and attribute that satisfy a notion of \n\\emph{compatibility}~\\cite{KoMaRaSc_JLC}. \nWe briefly discuss it for \\SAND~attack trees \nat the end of this section. \nWe start with a demonstration on how \nthe bottom-up evaluation \nalgorithm \ncan naturally be extended to \\SAND\\ attack trees.\n \n \nAn {\\em attribute domain for an attribute $\\fullattr$ on \\SAND\\ attack trees} \nis a tuple $\\attrdomain_\\attr = (\\attrval_\\attr, \\attror_\\attr, \\attrand_\\attr, \\attrsand_\\attr)$\nwhere $\\attrval_\\attr$ is a set of values and \n $\\attror_\\attr, \\attrand_\\attr, \\attrsand_\\attr$ are families of \n$k$-ary functions of the\n form $\\attrval_\\attr\\times\\dots\\times\\attrval_\\attr\\to\\attrval_\\attr$, \nassociated to \\OR, \\AND, and \\SAND\\ refinements, respectively. \nAn {\\em attribute for \\SAND~attack trees} is a pair \n$\\fullattr = (\\attrdomain_\\attr, \\basicassign_\\attr)$ \nformed by an attribute domain $\\attrdomain_\\attr$ and a function \n$\\basicassign_\\attr:\\basicact\\to\\attrval_\\attr$,\ncalled {\\em basic assignment} for $\\fullattr$, which associates a value from \n$\\attrval_\\attr$ with each basic action $b \\in \\basicact$. \n\\begin{definition}\n\\label{def:attr}\nLet $\\fullattr = \\big((\\attrval_\\attr, \\attror_\\attr, \\attrand_\\attr, \\attrsand_\n\\attr), \\basicassign_\\attr\\big)$\nbe an attribute. \nThe attribute evaluation function \n$\\attr: \\sandtree \\to \\attrval_\\attr$ which calculates the value of \nattribute \n$\\fullattr$ \nfor every \\SAND \\ attack tree $t \\in \\sandtree$\nis defined recursively as follows\n\\[ \\attr(t) = \\left\\{ \n \\begin{array}{l l}\n \\basicassign_\\attr(t) & \\quad \\text{if $t=b,\\ b\\in\\basicact$}\\\\\n \\attror_\\attr\\big(\\attr(t_1), \\dots, \\attr(t_k)\\big) & \\quad \\text{if $t=\\OR(t_1, \\dots, t_k)$}\\\\\n \\attrand_\\attr\\big(\\attr(t_1), \\dots, \\attr(t_k)\\big) & \\quad \\text{if $t=\\AND(t_1, \\dots, t_k)$}\\\\\n \\attrsand_\\attr\\big(\\attr(t_1), \\dots, \\attr(t_k)\\big) & \\quad \\text{if $t=\\SAND(t_1, \\dots, t_k)$}\n \\end{array} \\right.\\]\n\\end{definition}\n\n\nThe following example illustrates the bottom-up evaluation\nof the attribute \\emph{minimal attack time} on the \\SAND~attack trees \ngiven in Example~\\ref{eg:def-attack-tree}.\n\\begin{example}\n\\label{ex:min_cost}\nLet $\\attr$ denote the minimal time \nthat the attacker needs to achieve her goal.\nWe make the following assignments to the basic actions: \n$\\ftprhostsx\\mapsto 3$, $\\rshx\\mapsto 5$, $\\localbofx\\mapsto 7$, $\\sshbofx\\mapsto \n8$, $\\rsarefbofx\\mapsto 9$. \nSince we are interested in the minimal attack time, \nthe function for an \\OR\\ node \nis defined by \n$\\attror_\\attr(x_1,\\dots,x_k)=\\min\\{x_1,\\dots, x_k\\}$. \nThe function for an \\AND\\ node \nis $\\attrand_\\attr(x_1,\\dots,x_k)=\\max\\{x_1,\\dots, x_k\\}$, \nwhich models that the children \nof a conjunctively refined node are executed in parallel.\nFinally, in order to model that the children \nof a \\SAND\\ node need to be executed \nsequentially, \nwe let \n$\\attrsand_\\attr(x_1,\\dots,x_k)=\\sum_{i=1}^{k} x_i$. \nAccording to Definition~\\ref{def:attr}, \nthe minimal attack time for our running scenario $t$ is\n\\[\\attror_\\attr\\Big(\\attrsand_\\attr\\big(\\attrsand_\\attr(3,5),7\\big),\n\\attrand_\\attr(8, 9)\\Big)=\n\\min\\Big({\\mathrm\\Sigma}\\big({\\mathrm\\Sigma}(3,5),7\\big), \\max(8, 9)\\Big)=9.\\]\n\\end{example}\n\nIn the case of standard attack trees, \nthe bottom-up procedure uses only two functions to propagate \nthe attribute values to the root -- one for conjunctive and one for \ndisjunctive nodes. This means that the same function is employed \nto calculate the value of every conjunctively refined node, \nindependently of whether its children need to be executed sequentially or can \nbe executed simultaneously. \nEvidently, with \\SAND~attack trees, we can apply different propagation \nfunctions for \\AND\\ and \\SAND\\ nodes, as in \nExample~\\ref{ex:min_cost}. \nTherefore, \\SAND~attack trees can be evaluated over a larger set of \nattributes, and hence may provide more accurate evaluations of \nattack \nscenarios than \nstandard attack trees.\n\n\n\nTo guarantee that the evaluation of an attribute on equivalent \nattack trees yields\nthe same value, \nthe attribute domain must be \n\\emph{compatible} with a considered semantics~\\cite{KoMaRaSc_JLC}. \nOur complete set of axioms is a useful tool \nto check for compatibility. \nConsider an attribute domain\n$\\attrdomain_\\attr = (\\attrval_\\attr, \\attror_\\attr, \\attrand_\\attr, \\attrsand_\\attr)$, \nand let $\\sigma$ be a mapping\n$\\sigma=\\{\\OR\\mapsto \\attror_\\attr, \\AND \\mapsto \\attrand_\\attr, \n\\SAND \\mapsto \\attrsand_\\attr\\}$. \nGuaranteeing that $\\attrdomain_\\attr$ is compatible with \na semantics axiomatized by $E$ \namounts to verifying that the equality \n$\\sigma(l)=\\sigma(r)$ holds in $\\attrval_\\attr$, for every axiom\n$l=r\\in E$.\nIt is an easy exercise to show that the attribute domain \nfor minimal attack time, considered in Example~\\ref{ex:min_cost},\nis compatible with the SP semantics for \\SAND\\ attack trees.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Conclusions}\n\\label{sec:conclusions}\n\nWe have formalized the extension of attack trees \nwith sequential conjunctive refinement, \ncalled \\SAND, \nand given a semantics to \\SAND~attack trees in terms of\nsets of series-parallel graphs.\nThis SP semantics \nnaturally extends the multiset semantics \nfor attack trees from~\\cite{MaOo}.\nWe have shown that the notion of \na complete set of axioms for a semantics and \nthe bottom-up evaluation procedure can be generalized \nfrom attack trees to \\SAND~attack trees, \nand have proposed a complete axiomatization of the SP semantics. \n \n\nA number of recently proposed solutions\nfocus on extending attack trees with defensive\nmeasures~\\cite{RoKiTr2,KoMaRaSc_JLC}.\nThese extensions support reasoning about security scenarios\ninvolving two players -- an attacker and a defender --\nand the interaction between them.\nIn future work, we intend to add\nthe $\\SAND$ refinement to such trees.\nAfterwards, we plan to investigate\nsequential disjunctive refinement, as used\nfor instance in~\\cite{Arnold-POST14}. Our goal is to propose\na complete formalization\nof trees with attack and defense nodes,\nthat have parallel and sequential, conjunctive and disjunctive\nrefinements. \nThe findings will be implemented in the software \napplication ADTool~\\cite{adtool}. \n\n\n\n\\paragraph{\\textbf{Acknowledgments}} The research leading\nto these results has received funding from the European Union Seventh Framework\nProgramme under grant agreement number 318003 (TREsPASS) and from\nthe Fonds National de la Recherche Luxembourg under grant \nC13\/IS\/5809105.\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}