diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzziebz" "b/data_all_eng_slimpj/shuffled/split2/finalzziebz" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzziebz" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\nThe mathematical scattering theory for potential perturbations of the self-adjoint Laplacian in $L^{2}(\\mathbb{R}^{3})$ is a well developed subject. In particular, the stationary approach leads to a representation formula for the scattering matrix (see, e.g., \\cite{Y-LNM}). \nIn the case of singular perturbations, in particular, in the case of self-adjoint realizations of Laplacians with boundary conditions on hypersurfaces, similar results have been obtained in recent years, either using properties of the so-called Weyl function appearing in the resolvent formula (see \\cite{BMN}) or using a Limiting Absorption Principle (see \\cite{JST} and \\cite{JMPA}). The advantage of the approach by LAP over the former is that no trace class condition is required and so the results are not restricted to the two dimensional situation. \nThe target of the present paper is to put these separate results together and so to provide a representation formula for the scattering matrix in the case where the Laplacian is perturbed both by a regular short-range potential and by a singular one describing self-adjoint boundary conditions. In particular, we extend the abstract results in \\cite{JMPA} to a wider framework which then allows applications to such a setting. In order to develop such a strategy, at first we provide an abstract Kre\\u\\i n's type resolvent formula which, when applied to the concrete case of a potential perturbation of the Laplacian, allows for not compactly supported potentials. In particular, whenever the singular part of the perturbations is absent, this extends from compactly supported potentials in one dimension to short range potentials in three dimensions the kind of results provided in \\cite[Section 5]{BN}. \\par\nHere, in more details, the contents of the paper. In Section 2, following the scheme proposed in \\cite{JFA}, we provide an abstract resolvent formula for a perturbations $A_{\\mathsf B}$ of the self-adjoint $A$ by a linear combination of the adjoint of two bounded trace-like maps $\\tau_{1}:\\text{\\rm dom}(A)\\to\\mathfrak h_{1}$ and $\\tau_{2}:\\text{\\rm dom}(A)\\to\\mathfrak h_{2}$; while the kernel of $\\tau_{2}$ is required to be dense, so $\\tau_{2}^{*}$ plays the role of a singular perturbations, no further hypothesis is required for $\\tau_{1}$ and in applications that allows $\\tau_{1}^{*}$ to represent a regular perturbations by a short-range potential. In Subsection 2.3, by block operator matrices and the Schur complement, we re-write the obtained resolvent formula in terms of the resolvent of the operator corresponding to the non singular part of the perturbations; that plays an important role in the subsequent part regarding LAP and the scattering theory. In Section 3, following the scheme proposed in \\cite{JST} and further generalized in \\cite{JMPA}, at first we provide, under suitable hypothesis, a Limiting Absorption Principle for $A_{\\mathsf B}$ (see Theorem \\ref{LAP}) and then, by a combination of such a LAP with stationary scattering theory in the Birman-Yafaev scheme and the invariance principle, we obtain a representation formula for the scattering matrix of the couple $(A_{\\mathsf B},A)$ (see Theorem \\ref{S-matrix}). Whenever $A$ is the free Laplacian in $L^{2}(\\mathbb{R}^{3})$, such a formula contains, as subcases, both the usual formula for the perturbation given by a short-range potential as given, e.g., in \\cite{Y-LNM} and the formula for the case of a singular perturbation describing self-adjoint boundary conditions on a hypersurface as given in \\cite{JMPA}. Successively, in Section 4,\nin order to apply our abstract results to the case in which $A$ is the free Laplacian and the regular part represents a perturbation by a potential, we give various regularity results for the boundary layer operators associated to $\\Delta+\\mathsf v$, where $\\mathsf v$ is a potential of Kato-Rellich type. In the last Section we present various examples, where the free Laplacian is perturbed both by a singular term, describing either separating boundary conditions (as Dirichlet and Neumann ones) or semi-transparent (as $\\delta$ and $\\delta'$ type ones), and a regular one given by a short range potential $\\mathsf v$ decaying as $|x|^{-\\kappa(1+\\epsilon)}$, $\\epsilon>0$. In order to satisfy all our hypotheses, we need $\\kappa=2$. However, all our hypotheses but a single one (see Lemma \\ref{5.4}) hold with $\\kappa=1$; we conjecture that the requirement $\\kappa=2$ is merely of technical nature and that our results are true for a short range potential decaying as $|x|^{-(1+\\epsilon)}$. Finally, let us remark that whenever one is only interested in the construction of the operators and not in the scattering theory, then it is sufficient to assume that $\\mathsf v$ is a Kato-Rellich potential (see Remark \\ref{suff}). As a byproduct of our abstract construction (see Theorem \\ref{Th-alt-res}), the operator $A_{\\mathsf B}$ corresponds, whenever $A$ is the free Laplacian, to a singular perturbation of a Schr\\\"odinger operator. However, our concern here is the scattering theory with respect to the free Laplacian, thus we regard the regular and the singular parts of the perturbation as a single object; this constitutes the main novelty of our approach. Schr\\\"odinger operators with a Kato-Rellich potential plus a $\\delta$-like perturbation with a $p$-summable strength ($p>2$) have been already considered in \\cite{JDE18}, while for a different construction with a bounded potential and a $\\delta$- or a $\\delta'$-like perturbation with bounded strength we refer to \\cite{BLL}. None of such references considered the scattering matrix (however, \\cite{JDE18} provided a limiting absorption principle).\n\n\\subsection{Some notation and definition.}\n{\\ }\\par\n\\vskip5pt \\noindent $\\bullet$ $\\|\\cdot\\|_{X}$ denotes the norm on the complex Banach space $X$; in case $X$ is a Hilbert space, $\\langle\\cdot,\\cdot\\rangle_{X}$ denotes the (conjugate-linear w.r.t. the first argument) scalar product.\n\\vskip5pt\\noindent $\\bullet$ $\\langle\\cdot,\\cdot\\rangle_{X^{*},X}$ denotes the duality (assumed to be conjugate-linear w.r.t. the first argument) between the dual couple $(X^{*},X)$.\n\\vskip5pt\\noindent $\\bullet$ $L^{*}:\\text{\\rm dom}(L^{*})\\subseteq Y^{*}\\to X^{*}$ denotes the dual of the densely defined linear operator $L:\\text{\\rm dom}(L)\\subseteq X\\to Y$; in a Hilbert spaces setting $L^{*}$ denotes the adjoint operator.\n\\vskip5pt\\noindent $\\bullet$ $\\varrho(A)$ and $\\sigma(A)$ denote the resolvent set and the spectrum of the self-adjoint operator $A$; $\\sigma_{p}(A)$, $\\sigma_{pp}(A)$, $\\sigma_{ac}(A)$, $\\sigma_{sc}(A)$, denote the point, pure point, absolutely continuous and singular continuous spectra.\n\\vskip5pt\\noindent $\\bullet$ $\\mathscr B(X,Y)$, $\\mathscr B(X)\\equiv \\mathscr B(X,X)$, denote the Banach space of bounded linear operator on the Banach space $X$ to the Banach space $Y$; ${\\|}\\cdot {\\|}_{X,Y}$ denotes the corresponding norm.\n\\vskip5pt\\noindent $\\bullet$ ${\\mathfrak S}_{\\infty}(X,Y)$ denotes the space of compact operators on $X$ to $Y$.\n\\vskip5pt\\noindent $\\bullet$ $X\\hookrightarrow Y$ means that $X$ is continuously embedded into $Y$. \n\\vskip5pt\\noindent $\\bullet$ $\\Omega\\equiv\\Omega_{\\rm in}\\subset\\mathbb{R}^{3}$ denotes an open and bounded subset with a Lipschitz boundary $\\Gamma$; $\\Omega_{\\rm ex}:=\\mathbb{R}^{3}\\backslash\\overline\\Omega$.\n\\vskip5pt\\noindent $\\bullet$ $H^{s}(\\Omega)$ and $H^{s}(\\Omega_{\\rm ex})$ denote the scales of Sobolev spaces. \n\\vskip5pt\\noindent $\\bullet$ $H^{s}(\\mathbb{R}^{3}\\backslash\\Gamma):=H^{s}(\\Omega)\\oplus H^{s}(\\Omega_{\\rm ex})$.\n\\vskip5pt\\noindent $\\bullet$ $|x|$ denotes the norm of $x\\in\\mathbb{R}^{n}$.\n $\\langle x\\rangle$ denotes the function $x\\mapsto (1+|x|^{2})^{1\/2}$.\n\\vskip5pt\\noindent $\\bullet$ $L_{w}^{2}(\\mathbb{R}^{3})$, $w\\in\\mathbb{R}$, denotes the set of complex-valued functions $f$ such that $\\langle x\\rangle^{w}f\\in L^{2}(\\mathbb{R}^{3})$. \n\\vskip5pt\\noindent $\\bullet$ $H_{w}^{s}(\\mathbb{R}^{3}\\backslash\\Gamma):=H^{s}(\\Omega)\\oplus H_{w}^{s}(\\Omega_{\\rm ex})$, where $H_{w}^{s}(\\Omega_{\\rm ex})$ denotes the weighted Sobolev space relative to the weight $\\langle x\\rangle^{w}$.\n\\vskip5pt\\noindent $\\bullet$ $\\gamma_{0}^{\\rm in\/\\rm ex}$ and $\\gamma_{1}^{\\rm in\/\\rm ex}$ denote the interior\/exterior Dirichlet and Neumann traces on the boundary $\\Gamma$. \n\\vskip5pt\\noindent $\\bullet$ $\\gamma_{0}:=\\frac12(\\gamma_{0}^{\\rm in}+\\gamma_{0}^{\\rm ex})$, \n $\\gamma_{1}:=\\frac12(\\gamma_{1}^{\\rm in}+\\gamma_{1}^{\\rm ex})$.\n\\vskip5pt\\noindent $\\bullet$ $[\\gamma_{0}]:=\\gamma_{0}^{\\rm in}-\\gamma_{0}^{\\rm ex}$, \n $[\\gamma_{1}]:=\\gamma_{1}^{\\rm in}-\\gamma_{1}^{\\rm ex}$. \n\\vskip5pt\\noindent $\\bullet$ $S\\!L_{z}$ and $D\\!L_{z}$ denote the single- and double-layer operators. \n\\vskip5pt\\noindent $\\bullet$ $S_{z}:=\\gamma_{0}S\\!L_{z}$, $D_{z}:=\\gamma_{1}D\\!L_{z}$.\n\\vskip5pt\\noindent $\\bullet$ $D\\subset\\mathbb{R}$ is said to be discrete in the open set $E\\supset D $ whenever the (possibly empty) set of its accumulations point is contained in $\\mathbb{R}\\backslash E$; $D$ is said to be discrete whenever $E=\\mathbb{R}$.\n\\vskip5pt\\noindent $\\bullet$ Given $x\\ge 0$ and $y\\ge 0$, $x\\lesssim y$ means that there exists $c\\ge 0$ such that $x\\le c\\,y$.\n\n\\section{An abstract Kre\\u\\i n-type resolvent formula}\\label{Sec_Krein} \n\\subsection{The resolvent formula} Let $A:\\text{\\rm dom}(A)\\subseteq \\mathsf H\\to \\mathsf H$ be a self-adjoint operator in the Hilbert space $\\mathsf H$. We denote by $R_{z}:=(-A+z)^{-1}$, $z\\in \\varrho(A)$, its resolvent; one has $R_{z}\\in\\mathscr B(\\mathsf H,\\mathsf H_{A})$, where $\\mathsf H_{A}$ is the Hilbert space given by $\\text{\\rm dom} (A)$ equipped with the scalar product $$\\langle u,u\\rangle_{\\mathsf H_{A}}:=\\langle (A^{2}+1)^{1\/2} u,(A^{2}+1)^{1\/2}v\\rangle_{\\mathsf H}\\,.\n$$ \nLet\n$$\n\\mathfrak h_{k}\\hookrightarrow\\mathfrak h_{k}^{\\circ}\\hookrightarrow\\mathfrak h_{k}^{\\ast}\\,,\\qquad k=1,2\\,,\n$$\nbe auxiliary Hilbert spaces with dense continuous embedding; we do not identify $\\mathfrak h_{k}$ with its dual $\\mathfrak h^{*}_{k}$ (however, we use $\\mathfrak h_{k}\\equiv\\mathfrak h_{k}^{**}$) and we work with the $\\mathfrak h_{k}^{\\ast}$-$\\mathfrak h_{k}$ duality \n$\\langle\\cdot,\\cdot\\rangle_{\\mathfrak h_{k}^{\\ast},\\mathfrak h_{k}}$ defined in terms of the scalar\nproduct of the intermediate Hilbert space $\\mathfrak h_{k}^{\\circ}$. The scalar product and hence the duality are supposed to be conjugate linear with respect to the first variable; notice that $\\langle\\varphi,\\phi\\rangle_{\\mathfrak h_{k},\\mathfrak h^{\\ast}_{k}}=\\langle\\phi,\\varphi\\rangle^{*}_{\\mathfrak h_{k}^{\\ast},\\mathfrak h_{k}}$.\\par\nGiven the bounded linear maps\n$$\\tau_{k}:\\mathsf H_{A}\\rightarrow\\mathfrak h_{k}\\,,\\qquad k=1,2\\,,\n$$\nsuch that \n\\begin{equation}\\label{tau2}\n\\text{$\\text{\\rm ker}(\\tau_{2})$ is dense in $\\mathsf H$ and $\\text{\\rm ran}(\\tau_{2})$ is dense in $\\mathfrak h_{2}$,} \n\\end{equation}\nwe introduce the bounded operators \n$$\n\\tau:\\mathsf H_{A}\\to \\mathfrak h_{1}\\oplus\\mathfrak h_{2}\\,,\\qquad \\tau u:=\\tau_{1}u\\oplus\\tau_{2} u\\,,\n$$\nand\n$$\nG_{z}:\\mathfrak h_{1}^{\\ast}\\oplus \\mathfrak h_{2}^{*}\\to\\mathsf H\\,,\\qquad \nG_{z}:=(\\tau R_{\\bar{z}})^{\\ast}\\,,\\qquad z\\in\\varrho(A)\\,.\n$$\nWe further suppose that there exist reflexive Banach spaces $\\mathfrak b_{k}$, $k=1,2$, with dense continuous embeddings $\\mathfrak h_{k}\\hookrightarrow\\mathfrak b_{k}$ (hence $\\mathfrak b_{k}^{*}\\hookrightarrow\\mathfrak h_{k}^{*}$),\nsuch that\n$\\text{\\rm ran}(G_{z}|\\mathfrak b_{1}^{*}\\oplus\\mathfrak b_{2}^{*})$ is contained in the domain of definition of some (supposed to exist) $(\\mathfrak b_{1}\\oplus\\mathfrak b_{2})$-valued extension of $\\tau$ (which we denote by the same symbol) in such a way that \n\\begin{equation}\n\\tau G_{z}|\\mathfrak b_{1}^{*}\\oplus\\mathfrak b_{2}^{*}\\in{\\mathscr B}(\\mathfrak b_{1}^{*}\\oplus\\mathfrak b_{2}^{*},\\mathfrak b_{1}\\oplus\\mathfrak b_{2})\\,. \\label{tauG}%\n\\end{equation}\nGiven these hypotheses, \nwe set $\\mathsf B=(B_{0},B_{1},B_{2})$, with \n\\begin{equation}\\label{B012}\nB_{0}\\in{\\mathscr B}(\\mathfrak b_{2}^{*},\\mathfrak b_{2,2}^{*})\\,,\\quad B_{1}\\in\\mathscr B(\\mathfrak b_{1},\\mathfrak b_{1}^{*})\\,,\\quad\nB_{2}\\in\\mathscr B(\\mathfrak b_{2},\\mathfrak b^{*}_{2,2})\\,, \\quad\\text{$\\mathfrak b_{2,2}$ a reflexive Banach space,}\n\\end{equation}\n\\begin{equation}\\label{Bse}\n\\quad B_{1}=B_{1}^{*}\\,,\\qquad B_{0}B^{*}_{2}=B_{2}B^{*}_{0}\\,,\n\\end{equation}\nand introduce the map\n\\begin{equation}\\label{Lambda}\nZ_{\\mathsf B}\\ni z\\mapsto\\Lambda_{z}^{\\!\\mathsf B}\\in{\\mathscr B}(\\mathfrak b_{1}\\oplus\\mathfrak b_{2},\\mathfrak b_{1}^{*}\\oplus\\mathfrak b_{2}^{*})\n\\,,\\qquad \n\\Lambda_{z}^{\\!\\mathsf B}:=(M_{z}^{\\mathsf B})^{-1}( B_{1}\\oplus B_{2}) \\,,\n\\end{equation}\nwhere\n\\begin{equation}\\label{ZB}\nZ_{\\mathsf B}:=\\big\\{z\\in{\\mathbb{C}}\\backslash(-\\infty,0]: (M_{w}^{\\mathsf B})^{-1}\n\\in{\\mathscr B}(\\mathfrak b^{*}_{1}\\oplus\\mathfrak b^{*}_{2,2},\\mathfrak b^{*}_{1}\\oplus\\mathfrak b^{*}_{2})\\,,\\ w=z,\\bar{z}\\big\\}\n\\end{equation}\n$$\nM_{z}^{\\mathsf B}:=( 1\\oplus B_{0}) -( B_{1}\\oplus B_{2}) \\tau G_{z} \\in {\\mathscr B}(\\mathfrak b^{*}_{1}\\oplus\\mathfrak b^{*}_{2},\\mathfrak b^{*}_{1}\\oplus\\mathfrak b^{*}_{2,2})\\,.\n$$\n\\begin{theorem}\\label{Th_Krein} Suppose hypotheses \\eqref{tau2}, \\eqref{tauG}, \\eqref{B012} and \\eqref{Bse} hold and that $Z_{\\mathsf B}$ defined in \\eqref{ZB} is not empty. Then, defined $\\Lambda_{z}^{\\!\\mathsf B}$ as in \\eqref{Lambda},\n\\begin{equation}\nR_{z}^{\\mathsf B}:=R_{z}+G_{z}\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar{z}}^{\\ast}\\,,\\quad z\\in\nZ_{\\mathsf B}\\,, \\label{resolvent}%\n\\end{equation}\nis the resolvent of a self-adjoint operator $A_{\\mathsf B}$ and $Z_{\\mathsf B}%\n=\\varrho(A_{\\mathsf B})\\cap\\varrho(A)$.\n\\end{theorem}\n\\begin{proof} By \\eqref{Bse}, one gets\n\\begin{align*}\n\\big( ( 1\\oplus B_{0}) -( B_{1}\\oplus B_{2})\\tau\nG_{\\bar{z}}\\big)( B_{1}\\oplus B^{*}_{2}) =&( B_{1}\\oplus B_{2})\n\\big( ( 1\\oplus B_{0}^{*}) -\\tau G_{\\bar{z}}( B_{1}\\oplus B_{2}^{*})\\big)\\\\\n=&( B_{1}\\oplus B_{2})\n\\big( ( 1\\oplus B_{0}) -( B_{1}\\oplus B_{2})\\tau G_{{z}}\\big)^{*} \\,.\n\\end{align*}\nThis entails, by the definitions \\eqref{Lambda} and \\eqref{ZB},\n\\begin{equation}\\label{PS1}\n(\\Lambda^{\\!\\mathsf B}_{z})^{*}=\\Lambda^{\\!\\mathsf B}_{\\bar z}\\,.\n\\end{equation}\nBy the resolvent identity, there follows\n\\begin{align*}\n& (( 1\\oplus B_{0}) -( B_{1}\\oplus B_{2})\n\\tau G_{z}) -(( 1\\oplus B_{0}) -( B_{1}\\oplus B_{2}) \\tau G_{w}) \\\\ \n&=( B_{1}\\oplus B_{2}) \\tau(G_{w}-G_{z})\n=(z-w)( B_{1}\\oplus B_{2}) \\tau R_{w}G_{z}\\\\\n&=(z-w)(B_{1}\\oplus B_{2}) G_{\\bar{w}}^{\\ast}G_{z}\\,,\n\\end{align*}\nwhich entails \n\\begin{align*}\n& ( ( 1\\oplus B_{0}) -( B_{1}\\oplus B_{2})\n\\tau G_{w}) ^{-1}-(( 1\\oplus B_{0}) -\n(B_{1}\\oplus B_{2}) \\tau G_{z}) ^{-1} \\\\\n=&(z-w)(( 1\\oplus B_{0}) -( B_{1}\\oplus B_{2}) \\tau G_{w}) ^{-1}\n( B_{1}\\oplus B_{2})G_{\\bar{w}}^{\\ast}G_{z}(( 1\\oplus B_{0})\n -(B_{1}\\oplus B_{2}) \\tau G_{z}) ^{-1}\\,,\n\\end{align*}\nand hence\n\\begin{equation}\\label{PS2}\n\\Lambda_{w}^{\\!\\mathsf B}-\\Lambda_{z}^{\\!\\mathsf B}=(z-w)\\Lambda_{w}^{\\!\\mathsf B}G_{\\bar{w}}^{\\ast\n}G_{z}\\Lambda_{z}^{\\!\\mathsf B}\\,.\n\\end{equation}\nBy \\eqref{PS1} and $\\eqref{PS2}$, \n$$(R_{z}^{\\mathsf B})^{*}=R_{\\bar z}^{\\mathsf B}\\,,\\qquad R_{z}^{\\mathsf B}=R_{w}^{\\mathsf B}+(w-z)R_{z}^{\\mathsf B}R_{w}^{\\mathsf B}\\,.\n$$\n(see \\cite[page 113]{JFA}). Hence, $R_{z}^{\\mathsf B}$ is the resolvent of a self-adjoint operator whenever it is injective (see, e.g., \\cite[Theorems 4.10 and 4.19]{Stone}). By \\eqref{resolvent},\n\\begin{align*}\n&(B_{1}\\oplus B_{2})\\tau R_{z}^{\\mathsf B}=(B_{1}\\oplus B_{2})\\big(1+\\tau G_{z}\\Lambda_{z}^{\\!\\mathsf B}\\big)G_{\\bar z}^{*}=\\big((B_{1}\\oplus B_{2})+(B_{1}\\oplus B_{2})\\tau G_{z}\\Lambda_{z}^{\\!\\mathsf B}\\big)G_{\\bar z}^{*}\\\\\n=&\\big((B_{1}\\oplus B_{2})+\\big((1\\oplus B_{0})-\\big((1\\oplus B_{0})-(B_{1}\\oplus B_{2})\\tau G_{z}\\big)\\big)\\Lambda_{z}^{\\!\\mathsf B}\\big)G_{\\bar z}^{*}=(1\\oplus B_{0})\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar z}^{*}\\,.\n\\end{align*}\nThus, if $R_{z}^{\\mathsf B}u=0$ then \n$$\n0\\oplus 0=(1\\oplus B_{0})\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar z}^{*}u=\\big(\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar z}^{*}u\\big)_{1}\\oplus B_{0}\\big(\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar z}^{*}u\\big)_{2}\n$$\nBy\n$$\nG_{z}(\\phi_{1}\\oplus\\phi_{2})=G^{1}_{z}\\phi_{1}+G^{2}_{z}\\phi_{2}\\,, \\qquad G^{k}_{z}:=(\\tau_{k}R_{\\bar{z}})^{\\ast}\\,,\n$$\nthere follows \n\\begin{equation}\n0=R_{z}^{\\mathsf B}u=R_{z}u+G^{1}_{z}\\big(\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar{z}}^{\\ast}\nu\\big)_{1}+G^{2}_{z}\\big(\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar{z}}^{\\ast}u\\big)_{2}=R_{z}u+G^{2}_{z}\n\\big(\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar{z}}^{\\ast}u\\big)_{2}\\,.\\label{z2}%\n\\end{equation}\nSince the denseness of $\\text{\\rm ker}(\\tau_{2})$ is equivalent to $\\text{\\rm ker}(G^{2}_{z})=\\{0\\}$\nand the denseness of $\\text{\\textrm{ran}}(\\tau_{2})$ implies\n$\\text{\\textrm{ran}}(G^{2}_{z})\\cap \\text{\\rm dom}(A)=\\{0\\}$ \n(see \\cite[Remark 2.9]{JFA}), the relation \\eqref{z2} gives $\\big(\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar{z}}^{\\ast\n}u\\big)_{2}=0$. Thus $R_{z}^{\\mathsf B}u=0$ compels $R_{z}u=0$ and hence $u=0$.\\par \nFinally, the equality $Z_{\\mathsf B}=\\varrho(A_{\\mathsf B})\\cap\\varrho(A)$ is consequence of \\cite[Theorem 2.19 and\nRemark 2.20]{CFP}.\n\\end{proof}\n\\begin{remark} By \\eqref{resolvent}, if $u\\in\\text{\\rm dom}(A_{\\mathsf B})$, then $u=u_{0}+G_{z}(\\phi_{1}\\oplus\\phi_{2})$ for some $u_{0}\\in\\mathsf H_{A}$ and $\\phi_{1}\\oplus\\phi_{2}\\in \\mathfrak b_{1}^{*}\\oplus\\mathfrak b_{2}^{*}$; hence, by \\eqref{tauG}, \n$$\n\\tau:\\text{\\rm dom}(A_{\\mathsf B})\\to \\mathfrak b_{1}\\oplus\\mathfrak b_{2}\\,.\n$$ \n\\end{remark}\n\\subsection{An additive representation}\nAt first, let us introduce the Hilbert space $\\mathsf H_{A}^{*}$ defined as the completion of $\\mathsf H$ endowed with the scalar product $$\\langle u,v\\rangle_{\\mathsf H_{A}^{*}}:=\\langle(A^{2}+1)^{-1\/2}u,(A^{2}+1)^{-1\/2}v\\rangle_{\\mathsf H}\\,.\n$$\nNotice that that $R_{z}$ extends to a bounded bijective map (which we denote by the same symbol) on $\\mathsf H_{A}^{*}$ onto $\\mathsf H$. The linear operator $A$, being a densely defined bounded operator on $\\mathsf H$ to $\\mathsf H_{A}^{*}$, extends to a bounded operator $\\overline {\\!A}:\\mathsf H\\to\\mathsf H_{A}^{*}$ given by its closure. Moreover, denoting by $\\langle\\cdot,\\cdot\\rangle_{\\mathsf H_{A}^{*},\\mathsf H_{A}}$ the pairing obtained by extending the scalar product in $\\mathsf H$, since $A$ is self-adjoint and since $\\text{\\rm dom}(A)$ is dense in $\\mathsf H$,\n$$\n\\langle u,Av\\rangle_{\\mathsf H}=\\langle\\, \\overline{\\!A}u,v\\rangle_{\\mathsf H_{A}^{*},\\mathsf H_{A}}\\,,\\qquad u\\in\\mathsf H\\,,\\ v\\in\\mathsf H_{A}, \\,.\n$$\nFurther, we define $\\tau^{*}:\\mathfrak h_{1}^{*}\\oplus\\mathfrak h^{*}_{2}\\to\\mathsf H_{A}^{*}$ by \n\\begin{equation}\\label{tau*}\n\\langle \\tau^{*}\\phi,u\\rangle_{\\mathsf H_{A}^{*},\\mathsf H_{A}}=\\langle\\phi,\\tau u\\rangle_{\\mathfrak h_{1}^{*}\\oplus\\mathfrak h^{*}_{2},\\mathfrak h_{1}\\oplus\\mathfrak h_{2}}\\,,\\qquad u\\in\\mathsf H_{A}\\,,\\ \\phi\\in\\\\mathfrak h_{1}^{*}\\oplus\\mathfrak h^{*}_{2}\\,.\n\\end{equation}\nObviously, $\\tau^{*}(\\phi_{1}\\oplus\\phi_{2})=\\tau_{1}^{*}\\phi_{1}+\\tau^{*}_{2}\\phi_{2}$, where \n$\\tau_{k}^{*}:\\mathfrak h_{k}\\to\\mathsf H_{A}^{*}$, $k=1,2$, are defined in the same way as $\\tau^{*}$.\n\\par\nLet us notice that $R_{z}:\\mathsf H_{A}^{*}\\to \\mathsf H$ is the adjoint, with respect the pairing $\\langle\\cdot,\\cdot\\rangle_{\\mathsf H_{A}^{*},\\mathsf H_{A}}$, of $R_{\\bar z}:\\mathsf H_{A}\\to \\mathsf H$ and it is the inverse of $(-\\overline{\\!A} +z):\\mathsf H\\to\\mathsf H_{A}^{*}$; therefore \n\\begin{equation}\\label{blw}\nG_{z}=R_{z}\\tau^{*}\\,.\n\\end{equation}\n\\begin{lemma}\n\\label{le-green}Let $A_{\\mathsf B}:\\text{\\rm dom}(A_{\\mathsf B})\\subseteq\\mathsf H\\to\\mathsf H$ be the self-adjoint\noperator provided in Theorem \\ref{Th_Krein} and define\n\\begin{equation}\n\\rho_{\\mathsf B}:\\text{\\rm dom}(A_{\\mathsf B})\\rightarrow\\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*}\\,,\\quad\n\\rho_{\\mathsf B}(R_{z}^{\\mathsf B}u):=(\\pi_{1}^{\\ast}\\oplus 1)\\Lambda_{z}^{\\!\\mathsf B}G_{\\bar{z}}^{\\ast\n}u\\,,\\qquad u\\in \\mathsf H\\,, \\quad z\\in\\varrho(A_{\\mathsf B})\\cap\\varrho(A)\\,,\\label{rho_def}%\n\\end{equation}\nwhere $\\pi_{1}$ denotes the orthogonal projection onto the subspace\n$\\overline{\\text{\\rm ran}(\\tau_{1})}$. Then, the definition of $\\rho_{\\mathsf B}$ is\nwell-posed, i.e.,\n\\[\nR_{z_{1}}^{\\mathsf B}u_{1}=R_{z_{2}}^{\\mathsf B}u_{2}\\quad\\implies\\quad (\\pi_{1}^{\\ast}\\oplus 1)\\Lambda_{z_{1}}^{\\!\\mathsf B}G_{\\bar{z}_{1}}^{\\ast}u_{1}=(\\pi_{1}^{\\ast}\\oplus 1)\n\\Lambda_{z_{2}}^{\\!\\mathsf B}G_{\\bar{z}_{2}}^{\\ast}u_{2}%\n\\]\nand \n\\begin{equation}\\label{GF}\n\\langle u,\\Delta_{\\mathsf B}v\\rangle_{\\mathsf H}=\\langle A\nu,v\\rangle_{\\mathsf H}+\n\\langle\\tau u,\\rho_{\\mathsf B} v\\rangle_{\\mathfrak h_{1}\\oplus\\mathfrak h_{2},\\mathfrak h^{*}_{1}\\oplus\\mathfrak h_{2}^{*}}\\,,\\quad u\\in \\text{\\rm dom}(A),\\,v\\in\n\\text{\\rm dom}(A_{\\mathsf B})\\,.\n\\end{equation}\n\\end{lemma}\n\\begin{proof}\nLet $v=R_{z}^{\\mathsf B}u=v_{z}+G_{z}\\Lambda_{z}^{\\!\\mathsf B}\\tau v_{z}$, where $v_{z}%\n:=R_{z}u$ (hence $\\tau v_{z}=G_{\\bar{z}}^{\\ast}u$). Then\n\\begin{align*}\n& \\langle u,A_{\\mathsf B}v\\rangle_{\\mathsf H}-\\langle A\nu,v\\rangle_{\\mathsf H}\\\\\n= & -\\langle u,(-A_{\\mathsf B}+z)v\\rangle_{\\mathsf H}%\n+\\langle(-A+\\bar{z})u,v\\rangle_{\\mathsf H}\\\\\n= & -\\langle u,(-A+z)v_{z}\\rangle_{\\mathsf H}%\n+\\langle(-A+\\bar{z})u,v_{z}+G_{z}\\Lambda_{z}^{\\!\\mathsf B}\\tau v_{z}%\n\\rangle_{\\mathsf H}\\\\\n= & \\langle(-A+\\bar{z})u,G_{z}\\Lambda_{z}^{\\!\\mathsf B}\\tau v_{z}\\rangle\n_{\\mathsf H}=\\langle\\tau u,\\Lambda_{z}^{\\!\\mathsf B}\\tau v_{z}%\n\\rangle_{\\mathfrak h_{1}\\oplus\\mathfrak h_{2},\\mathfrak h^{*}_{1}\\oplus\\mathfrak h_{2}^{*}}\\\\\n= & \\langle (\\pi_{1}\\oplus 1)\\tau u,\\Lambda_{z}^{\\!\\mathsf B}\\tau v_{z}\\rangle\n_{\\mathfrak h_{1}\\oplus\\mathfrak h_{2},\\mathfrak h^{*}_{1}\\oplus\\mathfrak h_{2}^{*}}=\\langle\\tau u,(\\pi_{1}^{\\ast}\\oplus 1)\\Lambda\n_{z}^{\\!\\mathsf B}\\tau v_{z}\\rangle_{\\mathfrak h_{1}\\oplus\\mathfrak h_{2},\\mathfrak h^{*}_{1}\\oplus\\mathfrak h_{2}^{*}}\\,.\n\\end{align*}\nSuppose now that $R_{z_{1}}^{\\mathsf B}u_{1}=R_{z_{2}}^{\\mathsf B}u_{2}$. Then, by the above\nidentities, one gets, for any $u\\in \\text{\\rm dom}(A)$,\n\\[\n\\langle \\tau^{\\ast}(\\pi_{1}^{\\ast}\\oplus 1)(\\Lambda_{z_{1}}^{\\!\\mathsf B}G_{\\bar{z}_{1}%\n}^{\\ast}u_{1}-\\Lambda_{z_{2}}^{\\!\\mathsf B}G_{\\bar{z}_{2}}^{\\ast}u_{2}),u\\rangle\n_{\\mathsf H_{A}^{*},\\mathsf H_{A}}=0\\,.\n\\]\nHence $\\tau^{\\ast}((\\pi_{1}^{\\ast}\\oplus 1)\\Lambda_{z_{1}}^{\\!\\mathsf B}G_{\\bar{z}_{1}}%\n^{\\ast}u_{1}-(\\pi_{1}^{\\ast}\\oplus 1)\\Lambda_{z_{2}}^{\\!\\mathsf B}G_{\\bar{z}_{2}}^{\\ast}%\nu_{2})=0$. However, $\\text{\\rm ker}(\\tau^{\\ast})\\cap\\text{\\rm ran}((\\pi_{1}^{\\ast}\\oplus 1))=\\{0\\}$ since $\\pi_{1}^{\\ast}\\oplus 1$ is the projector onto the subspace\northogonal to $\\text{\\rm ker}(\\tau^{\\ast})$.\n\\end{proof}\nThe next Lemma provides a sort of abstract boundary conditions holding for the elements in $\\text{\\rm dom}(A_{\\mathsf B})$:\n\\begin{lemma}\\label{abc} Let $A_{\\mathsf B}$ be the self-adjoint operator in Theorem \\ref{Th_Krein}. Then, for any $z\\in\\varrho(A_{\\mathsf B})\\cap\\varrho(A)$, one has the representation\n$$\n\\text{\\rm dom}(A_{\\mathsf B})=\\{u\\in\\mathsf H:u_{z}:=u-G_{z}\\rho_{\\mathsf B}u\\in \\text{\\rm dom}(A)\\}\\,,\n$$\n$$\n(-A_{\\mathsf B}+z)u=(-A+z)u_{z}\\,.\n$$\nMoreover, \n$$\nu\\in \\text{\\rm dom}(A_{\\mathsf B})\\quad\\Longrightarrow\\quad (\\pi_{1}^{*}B_{1}\\oplus B_{2})\\tau u=(1\\oplus B_{0})\\rho_{\\mathsf B} u\\,.\n$$\n\\end{lemma}\n\\begin{proof} Since $G_{z}=R_{z}\\tau^{*}$ (see \\eqref{blw} below) and $\\pi_{1}^{*}\\oplus 1$ is the projection onto the orthogonal to $\\text{\\rm ker}(\\tau^{*})$, one has $G_{z}=G_{z}(\\pi_{1}^{*}\\oplus 1) $. \nHence, $u\\in\\text{\\rm dom}(A_{\\mathsf B})$ if and only if $u=R_{z}v+G_{z}(\\pi_{1}^{*}\\oplus 1) \\Lambda_{z}^{\\mathsf B}G^{*}_{\\bar z}v=R_{z}v+G_{z}\\rho_{\\mathsf B}u$. Therefore, \n$$\n\\text{\\rm dom}(A_{\\mathsf B})=\\{u\\in\\mathsf H:u=u_{z}+G_{z}\\rho_{\\mathsf B}u\\,, \\ u_{z}\\in\\text{\\rm dom}(A)\\}\\,.\n$$\nMoreover, given any $u\\in\\text{\\rm dom}(A)$, $u=R^{\\mathsf B}_{z}v$, one has\n$$\n(-A+z)u_{z}=(-A+z)R_{z}v=(-A_{\\mathsf B}+z)R^{\\mathsf B}_{z}v=(-A_{\\mathsf B}+z)u\\,.\n$$\nFinally, given $u=R^{\\mathsf B}_{z}v\\in\\text{\\rm dom} (A_{\\mathsf B})$, one has\n\\begin{align*}\n&(\\pi_{1}^{*}B_{1}\\oplus B_{2})\\tau u=(\\pi_{1}^{*}\\oplus 1)(B_{1}\\oplus B_{2})\\tau R^{\\mathsf B}_{z}v\\\\\n=&(\\pi_{1}^{*}\\oplus 1)\\big((B_{1}\\oplus B_{2})G_{\\bar z}v+ \n(B_{1}\\oplus B_{2})\\tau G_{z}\\big((1\\oplus B_{0})-(B_{1}\\oplus B_{2})\\tau G_{z}\\big)^{-1}(B_{1}\\oplus B_{2})G_{\\bar z}v\\big)\\\\\n=&(\\pi_{1}^{*}\\oplus 1)(1\\oplus B_{0})\\Lambda^{\\mathsf B}_{z}G_{\\bar z}v=(1\\oplus B_{0})(\\pi_{1}^{*}\\oplus 1)\\Lambda^{\\mathsf B}_{z}G_{\\bar z}v=(1\\oplus B_{0})\\rho_{\\mathsf B} u\\,.\n\\end{align*}\n\\end{proof} \nNow, we provide an additive representation of the self-adjoint $A_{\\mathsf B}$ in Theorem \\ref{Th_Krein}. \n\\begin{theorem}\\label{Th-add} Let $A_{\\mathsf B}:\\text{\\rm dom}(A_{\\mathsf B})\\subseteq \\mathsf H\\to\\mathsf H$ be the self-adjoint operator appearing in Theorem \\ref{Th_Krein}. Then\n$$\nA_{\\mathsf B}=\\overline {\\!A}+\\tau^{*}\\!\\rho_{\\mathsf B}\\,,\n$$\nwhere $\\rho_{\\mathsf B}$ is defined in \\eqref{rho_def}. In particular, if $B_{0}^{-1}\\in\\mathscr B(\\mathfrak b_{2,2}^{*},\\mathfrak b_{2}^{*})$, then\n$$\nA_{\\mathsf B}=\\overline {\\!A}+\\tau_{1}^{*}B_{1}\\tau_{1}+\\tau_{2}^{*}B_{0}^{-1}\\!B_{2}\\tau_{2}\\,.\n$$ \n\\end{theorem} \n\\begin{proof} By \\eqref{GF}, for any $u\\in\\text{\\rm dom}(A_{\\mathsf B})$ and $v\\in\\mathsf H_{A}$,\n\\begin{align*}\n\\langle \\Delta_{\\mathsf B}u,v\\rangle_{\\mathsf H_{A}^{*},\\mathsf H_{A}}\\equiv&\\langle \\Delta_{\\mathsf B}u,v\\rangle_{\\mathsf H}=\n\\langle u,Av\\rangle_{\\mathsf H}+\\langle\\rho_{\\mathsf B} u,\\tau v\\rangle_{\\mathfrak h^{*}_{1}\\oplus\\mathfrak h^{*}_{2},\\mathfrak h_{1}\\oplus\\mathfrak h_{2}}\\\\\n=&\\langle\\, \\overline{\\!A}u+\\tau^{*}\\!\\rho_{\\mathsf B} u,v\\rangle_{\\mathsf H_{A}^{*},\\mathsf H_{A}}\\,.\n\\end{align*}\nBy Lemma \\ref{abc} and by $\\tau_{1}^{*}\\pi_{1}^{*}=(\\pi_{1}\\tau_{1})^{*}=\\tau_{1}^{*}$,\n$$\n\\tau^{*}\\!\\rho_{\\mathsf B} =\\tau^{*}(\\pi^{*}_{1}B_{1}\\tau_{1}\\oplus B_{0}^{-1}B_{1}\\tau_{2})=\n\\tau_{1}^{*}B_{1}\\tau_{1}+\\tau_{2}^{*}B_{0}^{-1}\\!B_{2}\\tau_{2}\\,.\n$$\n\\end{proof}\n\n\n\\subsection{An alternative resolvent formula.} At first, let us notice that hypothesis \\eqref{tauG}, can be re-written as\n$$\n\\tau_{j}G^{k}_{z}|\\mathfrak b_{k}\\in\\mathscr B(\\mathfrak b_{k}^{*},\\mathfrak b_{j})\\,,\\quad j,k=1,2\\,,\\qquad G^{k}_{z}:=(\\tau_{k}R_{\\bar z})^{*}\\,.\n$$\nMoreover, \n$$\n{M}_{z}^{\\mathsf B}=(1\\oplus B_{0})+(B_{1}\\oplus B_{2})\\tau G_{z}=\n\\begin{bmatrix}M_{z}^{B_{1}}&B_{1}\\tau_{1}G^{2}_{z}\\\\\nB_{2}\\tau_{2}G^{1}_{z}&M_{z}^{B_{0},B_{2}}\n\\end{bmatrix}\n$$\nwhere\n$$\nM_{z}^{B_{1}}:=1-B_{1}\\tau_{1}G^{1}_{z}\\,,\\qquad \nM_{z}^{B_{0},B_{2}}:=B_{0}-B_{2}\\tau_{2}G^{2}_{z}\\,.\n$$\nThen, supposing all the inverse operators appearing in the next formula exist, by the inversion formula for block operator matrices, one gets\n\\begin{align}\\label{MB-1}\n&({M}_{z}^{\\mathsf B})^{-1}=\n\\begin{bmatrix}(M^{B_{1}}_{z})^{-1}+(M^{B_{1}}_{z})^{-1}B_{1}\\tau_{1}G^{2}_{z}({C}_{z}^{\\mathsf B})^{-1}B_{2}\\tau_{2}G^{1}_{z}(M^{B_{1}}_{z})^{-1}&\n(M^{B_{1}}_{z})^{-1}B_{1}\\tau_{1}G^{2}_{z}({C}_{z}^{\\mathsf B})^{-1}\\\\\n({C}_{z}^{\\mathsf B})^{-1}B_{2}\\tau_{2}G^{1}_{z}(M^{B_{1}}_{z})^{-1}&({C}_{z}^{\\mathsf B})^{-1}\n\\end{bmatrix}\n\\end{align}\nwhere ${C}^{\\mathsf B}_{z}$ denotes the second Schur complement, i.e.,\n\\begin{align*}\n{C}^{\\mathsf B}_{z}:=&M_{z}^{B_{0},B_{2}}-B_{2}\\tau_{2}G^{1}_{z}\n(M_{z}^{B_{1}})^{-1}B_{1}\\tau_{1}G^{2}_{z}\\\\\n=&M_{z}^{B_{0},B_{2}}\\left(1-(M_{z}^{B_{0},B_{2}})^{-1}B_{2}\\tau_{2}G^{1}_{z}\n(M_{z}^{B_{1}})^{-1}B_{1}\\tau_{1}G^{2}_{z}\\right)\\\\\n=&M_{z}^{B_{0},B_{2}}\\left(1-\\Lambda_{z}^{\\!B_{0},B_{2}}\\tau_{2}G^{1}_{z}\n\\Lambda_{z}^{\\!B_{1}}\\tau_{1}G^{2}_{z}\\right)\\,,\n\\end{align*}\n\\begin{equation}\\label{LB1}\n\\Lambda_{z}^{\\!B_{1}}:=(1-B_{1}\\tau_{1}G^{1}_{z})^{-1}B_{1}\\,,\n\\end{equation}\n\\begin{equation}\\label{LB02}\n\\Lambda_{z}^{\\!B_{0},B_{2}}:=(B_{0}-B_{2}\\tau_{2}G^{2}_{z})^{-1}B_{2}\\,.\n\\end{equation}\nRegarding the well-posedness of \\eqref{MB-1}, taking into account the definition of ${C}^{\\mathsf B}_{z}$, one has \n$$\nZ_{\\mathsf B}=\\big\\{z\\in\\varrho(A): (M^{\\mathsf B}_{z})^{-1}\\in{\\mathscr B}(\\mathfrak b^{*}_{1}\\oplus\\mathfrak b^{*}_{2,2},\\mathfrak b^{*}_{1}\\oplus\\mathfrak b^{*}_{2})\\,,\\ w=z,\\bar{z}\\big\\}\\supseteq \\widehat Z_{\\mathsf B}\\,,\n$$\nwhere\n\\begin{equation}\\label{wZB}\n\\widehat Z_{\\mathsf B}:=\\big\\{z\\in Z_{B_{1}}\\cap Z_{B_{0},B_{2}}:\\left(1-\\Lambda_{z}^{\\!B_{0},B_{2}}\\tau_{2}G^{1}_{w}\n\\Lambda_{z}^{\\!B_{1}}\\tau_{1}G^{2}_{w}\\right)^{-1}\\in\\mathscr B(\\mathfrak b_{2}^{*}),\\quad w=z,\\bar z\\big\\}\n\\end{equation}\n\\begin{equation}\\label{ZB1}\nZ_{B_{1}}:=\\big\\{z\\in\\varrho(A): (1-B_{1}\\tau_{1}G^{1}_{w})^{-1}\\in{\\mathscr B}(\\mathfrak b^{*}_{1})\\,,\\ w=z,\\bar{z}\\big\\}\\,,\n\\end{equation}\n\\begin{equation}\\label{ZB02}\nZ_{B_{0},B_{2}}:=\\big\\{z\\in\\varrho(A): (B_{0}-B_{2}\\tau_{2}G^{2}_{w})^{-1}\\in{\\mathscr B}(\\mathfrak b^{*}_{2,2},\\mathfrak b^{*}_{2})\\,,\\ w=z,\\bar{z}\\big\\}\\,,\n\\end{equation}\nTherefore, supposing that $\\widehat Z_{\\mathsf B}$ is not empty, for any $z\\in\\widehat Z_{\\mathsf B}$, by \\eqref{resolvent} and by\n$$\n({C}_{z}^{\\mathsf B})^{-1}B_{2}=\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\\,,\\qquad \\Sigma^{\\mathsf B}_{z}:=\\left(1-\\Lambda_{z}^{\\!B_{0},B_{2}}\\tau_{2}G^{1}_{z}\n\\Lambda_{z}^{\\!B_{1}}\\tau_{1}G^{2}_{z}\\right)^{-1}\\,,\n$$\n one has\n\\begin{align*}\n\\Lambda_{z}^{\\!\\mathsf B}=(M^{\\mathsf B}_{z})^{-1}\n\\begin{bmatrix}B_{1}&0\\\\\n0&B_{2}\\end{bmatrix}\n=&\n\\begin{bmatrix}\\Lambda^{\\!B_{1}}_{z}+\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}&\n\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\\\\\n\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}&\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\n\\end{bmatrix}\n\\,.\n\\end{align*}\nTherefore\n\\begin{equation}\\label{res}\nR_{z}^{\\mathsf B}=R_{z}+\\begin{bmatrix}G^{1}_{z}&G^{2}_{z}\\end{bmatrix}\n\\begin{bmatrix}\\Lambda^{\\!B_{1}}_{z}+\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}&\n\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\\\\\n\\Sigma^{\\mathsf B}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}&\\Sigma^{\\mathsf B}_{z}\n\\Lambda_{z}^{\\!B_{0},B_{2}}\\end{bmatrix}\n\\begin{bmatrix}{G^{1*}_{\\bar z}}\\\\{G^{2*}_{\\bar z}}\\end{bmatrix}\\,.\n\\end{equation}\nIn particular, taking $\\mathsf B=(1,B_{1},0)$, one gets, for any $z\\in Z_{B_{1}}$,\n\\begin{align}\\label{res1}\nR_{z}^{B_{1}}:=R_{z}^{(1,B_{1},0)}\n=&R_{z}+\\begin{bmatrix}G^{1}_{z}&G^{2}_{z}\\end{bmatrix}\n\\begin{bmatrix}\\Lambda_{z}^{\\!B_{1}}&0\\\\\n0&0\n\\end{bmatrix}\n\\begin{bmatrix}{G^{1*}_{\\bar z}}\\\\{G^{2*}_{\\bar z}}\\end{bmatrix}\n=R_{z}+G^{1}_{z}\\Lambda_{z}^{\\!B_{1}}{G^{1*}_{\\bar z}}\n\\end{align}\nwhile, taking $\\mathsf B=(B_{0},0,B_{2})$, one gets, for any $z\\in Z_{B_{0},B_{2}}$, \n\\begin{align}\\label{res2}\nR_{z}^{B_{0},B_{2}}:=R_{z}^{(B_{0},0,B_{2})}\n=&R_{z}+\\begin{bmatrix}G^{1}_{z}&G^{2}_{z}\\end{bmatrix}\n\\begin{bmatrix}0&\n0\\\\\n0&\\Lambda_{z}^{\\!B_{0},B_{2}}\n\\end{bmatrix}\n\\begin{bmatrix}{G^{1*}_{\\bar z}}\\\\{G^{2*}_{\\bar z}}\\end{bmatrix}\n=R_{z}+G^{2}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}{G^{2*}_{\\bar z}}\\,.\n\\end{align}\nTherefore, by Theorem \\ref{Th_Krein} with $\\mathsf B=(B_{1},0,0)$, one gets\n\\begin{corollary}\\label{cor1} Let $\\tau_{1}\\in\\mathscr B(\\mathsf H_{A},\\mathfrak h_{1})$ such that $\\tau_{1}G^{1}_{z}|\\mathfrak b^{*}_{1}\\in\\mathscr B(\\mathfrak b_{1}^{*},\\mathfrak b_{1})$ and let $B_{1}\\in\\mathscr B(\\mathfrak b_{1},\\mathfrak b^{*}_{1})$ self-adjoint; suppose that \n$Z_{B_{1}}$ defined in \\eqref{ZB1} is not empty. Then \n\\begin{equation}\\label{res1.1}\nR_{z}^{B_{1}}=R_{z}+G^{1}_{z}\\Lambda_{z}^{\\!B_{1}}{G^{1*}_{\\bar z}}\\,,\\qquad z\\in Z_{B_{1}}\\,,\n\\end{equation}\nwhere $\\Lambda_{z}^{\\!B_{1}}$ is defined in \\eqref{LB1}, is the resolvent of a self-adjoint operator $A_{B_{1}}$ and $Z_{B_{1}}=\\varrho(A_{B_{1}})\\cap\\varrho(A)$.\n\\end{corollary}\nBy Theorem \\ref{Th_Krein} with $\\mathsf B=(B_{0},0,B_{2})$, one gets\n\\begin{corollary}\\label{cor02} Let $\\tau_{2}\\in\\mathscr B(\\mathsf H_{A},\\mathfrak h_{2})$ satisfy \\eqref{tau2} be such that $\\tau_{1}G^{1}_{z}|\\mathfrak b^{*}_{2}\\in\\mathscr B(\\mathfrak b_{2}^{*},\\mathfrak b_{2})$ and let $B_{0}\\in\\mathscr B(\\mathfrak b_{2}^{*},\\mathfrak b_{2,2}^{*})$, $B_{2}\\in \\mathscr B(\\mathfrak b_{2},\\mathfrak b_{2,2}^{*})$ be such that $B_{0}B_{2}^{*}=B_{2}B_{0}^{*}$; suppose that\n$Z_{B_{0},B_{2}}$ defined in \\eqref{ZB02} is not empty. Then \n\\begin{equation}\\label{res2.1}\nR_{z}^{B_{0},B_{2}}=R_{z}+G^{2}_{z}\\Lambda_{z}^{\\!B_{0},B_{2}}{G^{2*}_{\\bar z}}\\,,\\qquad z\\in Z_{B_{0},B_{2}}\n\\end{equation}\nwhere $\\Lambda_{z}^{\\!B_{0},B_{2}}$ is defined in \\eqref{LB02}, is the resolvent of a self-adjoint operator $A_{B_{0},B_{2}}$ and $Z_{B_{0},B_{2}}=\\varrho(A_{B_{0},B_{2}})\\cap\\varrho(A)$.\n\\end{corollary}\nSupposing $\\widehat Z_{\\mathsf B}\\not=\\varnothing $, by \\eqref{res}, by \\eqref{res1} and by the relations\n\\begin{align}\\label{GB1}\nG^{B_{1}}_{z}:=(\\tau_{2}R_{\\bar z}^{B_{1}})^{*}=&\n(\\tau_{2}R_{\\bar z}+\\tau_{2} G^{1}_{\\bar z}\\Lambda^{\\!B_{1}}_{\\bar z}{G^{1*}_{z}})^{*}\\\\\n=&G_{z}^{2}+G_{z}^{1}\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G_{z}^{2}\\nonumber\n\\end{align}\n\\begin{align*}\n{G^{B_{1}*}_{\\bar z}}=\\tau_{2}R_{z}^{B_{1}}=&\n\\tau_{2}R_{z}+\\tau_{2} G^{1}_{ z}\\Lambda^{\\!B_{1}}_{z}{G^{1*}_{\\bar z}}\\\\=&G_{\\bar z}^{2*}+\\tau_{2}G_{z}^{1}\\Lambda^{\\!B_{1}}_{z}G_{\\bar z}^{1*}\n\\end{align*}\n\\begin{align*}\n\\widehat M_{z}^{\\mathsf B}=B_{0}-B_{2}\\tau_{2}G^{B_{1}}_{z}=&B_{0}-B_{2}\\tau_{2}G_{z}^{2}+\\tau_{2}G_{z}^{1}\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G_{z}^{2}\\\\\n=&M_{z}^{B_{0},B_{2}}+B_{2}\\tau_{2}G_{z}^{1}\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G_{z}^{2}\\\\\n=&M_{z}^{B_{0},B_{2}}\\big(1+\\Lambda^{\\!B_{0},B_{1}}_{z}\\tau_{2}G_{z}^{1}\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G_{z}^{2}\\big)\n\\end{align*}\n\\begin{equation}\\label{wLB1}\n\\widehat \\Lambda^{\\mathsf B}_{z}:=(\\widehat M_{z}^{\\mathsf B})^{-1}B_{2}=(B_{0}-B_{2}\\tau_{2}G^{B_{1}}_{z})^{-1}B_{2}=\\Sigma_{z}^{\\mathsf B}\\Lambda^{\\!B_{0},B_{2}}_{z}\n\\end{equation}\n one gets \n\\begin{align}\n\\Lambda_{z}^{\\mathsf B}=&\\begin{bmatrix}\\Lambda^{\\!B_{1}}_{z}+\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}&\n\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\\\\n\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda_{z}^{\\!B_{1}}&\n\\widehat \\Lambda^{\\mathsf B}_{z}\\,;\n\\end{bmatrix}\\label{LB-new}\\\\\n=&\\left(1+\\begin{bmatrix}\\Lambda^{\\!B_{1}}_{z}&0\\\\0&\\widehat \\Lambda^{\\mathsf B}_{z}\n\\end{bmatrix}\\begin{bmatrix}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}&\n\\tau_{1}G^{2}_{z}\\\\\n\\tau_{2}G^{1}_{z}&0\n\\end{bmatrix}\\,\\right)\\begin{bmatrix}\\Lambda^{\\!B_{1}}_{z}&0\\\\0&\\widehat \\Lambda^{\\mathsf B}_{z}\n\\end{bmatrix}\\,.\\label{LB-new2}\n\\end{align}\nTherefore \n\\begin{align}\\label{res-new}\nR_{z}^{\\mathsf B}=&R_{z}+\\begin{bmatrix}G^{1}_{z}&G^{2}_{z}\\end{bmatrix}\n\\begin{bmatrix}\\Lambda^{\\!B_{1}}_{z}+\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}&\n\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\\\\n\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda_{z}^{\\!B_{1}}&\n\\widehat \\Lambda^{\\mathsf B}_{z}\n\\end{bmatrix}\n\\begin{bmatrix}\n{G^{1*}_{\\bar z}}\\\\ \n{G^{2*}_{\\bar z}}\n\\end{bmatrix}\\\\\n=&R_{z}+\\begin{bmatrix}G^{1}_{z}&G^{2}_{z}\\end{bmatrix}\n\\begin{bmatrix}\\Lambda^{\\!B_{1}}_{z}G^{1*}_{\\bar z}+\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}G^{1*}_{\\bar z}+\n\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}G^{2*}_{\\bar z}\\\\\n\\widehat \\Lambda^{\\!\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda_{z}^{\\!B_{1}}G^{1*}_{\\bar z}+\n\\widehat \\Lambda^{\\mathsf B}_{z}G^{2*}_{\\bar z}\n\\end{bmatrix}\\nonumber\\\\\n=&R_{z}+G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}G^{1*}_{\\bar z}+G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda^{\\!B_{1}}_{z}G^{1*}_{\\bar z}+\nG^{1}_{z}\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}G^{2*}_{\\bar z}\\nonumber\\\\\n&+G^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}\\tau_{2}G^{1}_{z}\\Lambda_{z}^{\\!B_{1}}G^{1*}_{\\bar z}+\nG^{2}_{z}\\widehat \\Lambda^{\\mathsf B}_{z}G^{2*}_{\\bar z}\\nonumber\\\\\n=&R^{B_{1}}_{z}+G_{z}^{B_{1}}\\widehat \\Lambda^{\\mathsf B}_{z}G_{\\bar z}^{B_{1}*}\\nonumber\\,.\n\\end{align}\nThis also entails, by \\cite[Theorem 2.19 and Remark 2.20]{CFP}, that if $\\widehat Z_{\\mathsf B}\\not=\\varnothing $, then $\\widehat Z_{\\mathsf B}=Z_{\\mathsf B}=\\varrho(A_{\\mathsf B})\\cap\\varrho(A_{B_{1}})$. Summing up, one has the following \n\\begin{theorem}\\label{Th-alt-res} Assume that hypotheses \\eqref{tauG}, \\eqref{B012} and \\eqref{Bse} hold and that $\\widehat Z_{\\mathsf B}$ defined in \\eqref{wZB} is not empty. Then, for any $z\\in \\varrho(A_{\\mathsf B})\\cap\\varrho(A_{B_{1}})$, the resolvent $R_{z}^{\\mathsf B}$ in \\eqref{resolvent} has the representation \\eqref{res-new} and \n\\begin{equation}\\label{alt-res}\nR^{\\mathsf B}_{z}=R_{z}^{B_{1}}+G_{z}^{B_{1}}\\widehat \\Lambda^{\\mathsf B}_{z}G_{\\bar z}^{B_{1}*}\\,,\\qquad\nz\\in \\varrho(A_{\\mathsf B})\\cap\\varrho(A_{B_{1}})\\,,\n\\end{equation}\nwhere $R_{z}^{B_{1}}$, $G_{z}^{B_{1}}$ and $\\widehat \\Lambda^{\\mathsf B}_{z}$ are defined in \\eqref{res1.1}, \\eqref{GB1} and \\eqref{LB1}.\n\\end{theorem}\n\\begin{remark}Let us notice that the resolvent formula \\eqref{alt-res} is of the same kind of the one in \\eqref{res2.1}, whenever one replaces $A$ with $A_{B_{1}}$. \n\\end{remark}\nBy using the same kind of arguments as in the proof of Lemma \\ref{abc} and defining \n$$\n\\widehat \\rho_{\\mathsf B}:\\text{\\rm dom}(A_{\\mathsf B})\\to\\mathfrak h_{2}^{*}\\,,\\qquad \\widehat \\rho_{\\mathsf B}(R^{\\mathsf B}_{z}u):=G_{z}^{B_{1}}\\widehat\\Lambda_{z}^{\\mathsf B}G^{B_{1}*}_{\\bar z}u\\,,\n$$\none gets the following\n\\begin{lemma}\\label{alt-abc} Let $A_{\\mathsf B}$ be the self-adjoint operator in Theorem \\ref{Th-alt-res}. Then, for any $z\\in\\varrho(A_{\\mathsf B})\\cap\\varrho(A_{B_{1}})$, one has the representation\n$$\n\\text{\\rm dom}(A_{\\mathsf B})=\\{u\\in\\mathsf H:u_{z}:=u-G^{B_{1}}_{z}\\widehat\\rho_{\\mathsf B}u\\in \\text{\\rm dom}(A_{B_{1}})\\}\\,,\n$$\n$$\n(-A_{\\mathsf B}+z)u=(-A_{B_{1}}+z)u_{z}\\,.\n$$\nMoreover, \n$$\nu\\in \\text{\\rm dom}(A_{\\mathsf B})\\quad\\Longrightarrow\\quad B_{2}\\tau_{2} u=B_{0}\\widehat\\rho_{\\mathsf B} u\\,.\n$$\n\\end{lemma}\n\n\\section{The Limiting Absorption Principle and the Scattering Matrix}\\label{Sec-LAP}\nNow we suppose that $\\mathsf H=L^{2}(M,{\\mathcal B},m)\\equiv L^{2}(M)$. \nGiven a measurable $\\varphi:M\\to [1,+\\infty)$, we define the weighted $L^{2}$-space \n\\begin{equation}\\label{Lphi}\nL_{\\varphi}^{2}(M,{\\mathcal B},m)\\equiv L_{\\varphi}^{2}(M):=\\{\\text{$u:M\\to\\mathbb{C}$ measurable} : \\varphi u\\in L^{2}(M)\\}\\,.\n\\end{equation}\nBy $\\varphi\\ge 1$, \n$$\nL^{2}_{\\varphi}(M)\\hookrightarrow L^{2}(M)\\hookrightarrow L_{\\varphi^{-1}}^{2}(M)\\simeq L_{\\varphi}^{2}(M)^{*}\\,.\n$$\nFrom now on $\\langle\\cdot,\\cdot\\rangle $ and $\\|\\cdot\\|$ denote the scalar product and the corresponding norm on $L^{2}(M)$; $\\langle\\cdot,\\cdot\\rangle_{\\varphi} $ and $\\|\\cdot\\|_{\\varphi}$ denote the scalar product and the corresponding norm on $L_{\\varphi}^{2}(M)$. \\par\nThen we introduce the following hypotheses: \\vskip5pt\\par\\noindent\n(H1) $A_{B_{1}}$ is bounded from above and there exists a positive $\\lambda_{1}\\ge\\sup\\sigma(A_{B_{1}})$, such that \n$R^{B_{1}}_{z}\\in \\mathscr B(L^{2}_{\\varphi}(M))$ for any $z\\in \\varrho(A_{B_{1}})$ such that $\\text{Re}(z)> \\lambda_{1}$; \n\\vskip5pt\\par\\noindent\n(H2) $A_{B_{1}}$ satisfies a Limiting Absorption Principle (LAP for short), i.e. there exists a (eventually empty) closed set with zero Lebesgue measure ${e}(A_{B_{1}})\\subset \\mathbb{R}$ such that, for all $\\lambda\\in \\mathbb{R}\\backslash{e}(A_{B_{1}})$, the limits\n\\begin{equation}\\label{LAP1}\nR^{B_{1},\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}R^{B_{1}}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(L_{\\varphi}^{2}(M),L_{\\varphi^{-1}}^{2}(M))$ and the maps $z\\mapsto R^{B_{1},\\pm}_{z}$, where $R^{B_{1},\\pm}_{z}\\equiv R^{B_{1}}_{z}$ whenever $z\\in\\varrho(A_{B_{1}})$, are continuous on $(\\mathbb{R}\\backslash{e}(A_{B_{1}}))\\cup\\mathbb{C}_{\\pm}$ to $\\mathscr B(L_{\\varphi}^{2}(M),L_{\\varphi^{-1}}^{2}(M))$;\n\\vskip5pt\\par\\noindent\n(H3) for any compact set $K\\subset \\mathbb{R}\\backslash{e}(A_{B_{1}})$ there exists $c_{K}>0$ such that for any $\\lambda\\in K$ and for any $u\\in L_{\\varphi^{2}}^{2}(M)\\cap\\text{\\rm ker}(R^{B_{1},+}_{\\lambda}-R^{B_{1},-}_{\\lambda})$ one has\n\\begin{equation}\\label{(H3)}\n\\|R^{B_{1},\\pm}_{\\lambda}u\\|\\le c_{K}\\|u\\|_{\\varphi^{2}}\\,;\n\\end{equation} \nWe split next hypothesis (H4) in two separate points:\\vskip5pt\\par\\noindent\n(H4.1) $A_{\\mathsf B}$ is bounded from above;\n\\vskip5pt\\par\\noindent\n(H4.2) the embedding $\\mathfrak h_{2}\\hookrightarrow\\mathfrak b_{2}$ is compact and there exists a positive $\\lambda_{2}>\\sup\\sigma(A_{B_{1}})$, such that $G_{z}^{B_{1}}\\in \\mathscr B(\\mathfrak h^{*}_{2},L_{\\varphi^{2+\\gamma}}^{2}(M))$ for some $\\gamma>0$ and for any $z\\in\\varrho(A_{B_{1}})$ such that $\\text{Re}(z)>\\lambda_{2}$.\n\\vskip5pt\\par\\noindent\nThen, $A_{\\mathsf B}$ satisfies a Limiting Absorption Principle as well:\n\\begin{theorem}\\label{LAP} Suppose hypotheses (H1)-(H4) hold. Then the limits\n\\begin{equation}\\label{LAP2}\nR^{\\mathsf B,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}R^{\\mathsf B}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(L_{\\varphi}^{2}(M),L_{\\varphi^{-1}}^{2}(M))$ for all $\\lambda\\in\\mathbb{R}\\backslash{e}(A_{\\mathsf B})$, where ${e}(A_{\\mathsf B}):={e}(A_{B_{1}})\\cup\\sigma_{p}(A_{\\mathsf B})$, and \n${e}(A_{\\mathsf B})\\backslash{e}(A_{B_{1}})$ is a (possibly empty) discrete set in $\\mathbb{R}\\backslash {e}(A_{B_{1}})$; the maps $z\\mapsto R^{\\mathsf B,\\pm}_{z}$, where $R^{\\mathsf B,\\pm}_{z}\\equiv R^{\\mathsf B}_{z}$ whenever $z\\in\\varrho(A_{\\mathsf B})$, are continuous on $(\\mathbb{R}\\backslash {e}(A_{\\mathsf B}))\\cup\\mathbb{C}_{\\pm}$ to $\\mathscr B(L_{\\varphi}^{2}(M),L_{\\varphi^{-1}}^{2}(M))$\n\\end{theorem}\n\\begin{proof} We use \\cite[Theorem 3.1]{JMPA} (which builds on \\cite{Reng}). By (H1), \\eqref{alt-res} and (H4.2), $R^{B_{1}}_{z}$ and $R^{\\mathsf B}_{z}$ are in $\\mathscr B(L^{2}_{\\varphi}(M))$ and $z\\mapsto R^{B_{1}}_{z}$ and $z\\mapsto R^{\\mathsf B}_{z}$ are continuous since pseudo-resolvents in $\\mathscr B(L^{2}_{\\varphi}(M))$; $A_{\\mathsf B}$ is bounded from above by (H4.1). Therefore hypothesis (H1) in \\cite{JMPA} holds true. Our hypotheses (H2) and (H3) coincides with the same ones in \\cite{JMPA}. By (H4.2), the embedding $\\mathfrak b^{*}_{2}\\hookrightarrow\\mathfrak h^{*}_{2}$ is compact. From $\\widehat\\Lambda^{\\mathsf B}_{z}\\in\\mathscr B(\\mathfrak b_{2},\\mathfrak b_{2}^{*})$ and \\eqref{alt-res}, follows that \n$R^{\\mathsf B}_{z}-R^{B_{1}}_{z}\\in{\\mathfrak S}_{\\infty}(L^{2}(M),L_{\\varphi^{2+\\gamma}}^{2}(M))$. Therefore hypothesis (H4) in \\cite{JMPA} holds and the statement is a consequence of \\cite[Theorem 3.1]{JMPA}.\n\\end{proof}\nLet us now assume that \\vskip8pt\\noindent\n(H5) the limits\n\\begin{equation}\\label{limG}\nG_{\\lambda}^{B_{1},\\pm}:=\\lim_{\\epsilon\\searrow 0}G^{B_{1}}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(\\mathfrak h_{2} ^{*},L^{2}_{\\varphi^{-1}}(M))$ for any $\\lambda\\in\\mathbb{R}\\backslash{e}(A_{B_{1}})$ and the maps $z\\mapsto G_{z}^{B_{1},\\pm}$, where $G_{z}^{B_{1},\\pm}\\equiv G^{B_{1}}_{z}$ whenever $z\\in\\varrho(A_{B_{1}})$, are continuous on $(\\mathbb{R}\\backslash{e}(A_{B_{1}}))\\cup\\mathbb{C}_{\\pm}$ to $\\mathscr B(\\mathfrak h_{2} ^{*},L^{2}_{\\varphi^{-1}}(M))$; moreover, the linear operators $G_{z}^{B_{1},\\pm}$ are injective. \n\\vskip8pt\\noindent\nThen, by \\cite[Lemma 3.6]{JMPA}, one gets the following:\n\\begin{lemma}\\label{bound}\nAssume that (H1)-(H5) hold. Then, for any open and bounded $I$ s.t. $\\overline{I}\\subset \\mathbb{R}\\backslash{e}(A_{\\mathsf B})$, one has \n\\begin{equation}\\label{supLambda}\n\\sup_{(\\lambda,\\epsilon)\\in I\\times (0,1)}{\\|}\\widehat\\Lambda^{\\mathsf B}_{\\lambda\\pm i\\epsilon}{\\|}_{\\mathfrak h_{2} ,\\mathfrak h_{2}^{*}}<+\\infty\\,.\n\\end{equation}\nMoreover, for any $\\lambda\\in \\mathbb{R}\\backslash{e}(A_{\\mathsf B})$, the limits\n\\begin{equation}\\label{limLambda}\n\\widehat\\Lambda^{\\mathsf B,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\widehat\\Lambda^{\\mathsf B}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(\\mathfrak h_{2} ,\\mathfrak h_{2} ^{*})$ and\n\\begin{equation}\\label{limRes}\nR^{\\mathsf B,\\pm}_{\\lambda}=R^{B_{1},\\pm}_{\\lambda}+G^{B_{1},\\pm}_{\\lambda}\\widehat\\Lambda^{\\mathsf B,\\pm}_{\\lambda}(G^{B_{1},\\mp}_{\\lambda})^{*}\\,.\n\\end{equation}\n\\end{lemma}\nBy the same reasoning as at the end of \\cite[proof of Theorem 5.1]{JMPA}, one can improve the result regarding \\eqref{limLambda}: \n\\begin{corollary}\\label{limbb} Suppose hypotheses (H1)-(H5) hold. Then the limits \\eqref{limLambda} exist in $\\mathscr B(\\mathfrak b_{2},\\mathfrak b_{2}^{*})$.\n\\end{corollary}\nBefore stating the next results, we recall the following: \n\\begin{definition} Given two self-adjoint operators $A_{1}$ and $A_{2}$ in the Hilbert space $\\mathsf H$, we say that completeness holds for the scattering couple $(A_{1},A_{2})$ whenever the strong limits \n$$\nW_{\\pm}(A_ {1},A_{2})\n:=\\text{s-}\\lim_{t\\to\\pm\\infty}e^{itA_{1}}e^{-itA_{2}}P_{2}^{ac}\n\\,,\\qquad\nW_{\\pm}(A_{2},A_{1})\n:=\\text{s-}\\lim_{t\\to\\pm\\infty}e^{itA_{2}}e^{-itA_{1}}P_{1}^{ac}\\,,\n$$ \nexist everywhere in $\\mathsf H$ and \n$$\\text{\\rm ran}(W_{\\pm}(A_ {1},A_{2}))=\\mathsf H_{1}^{ac }\\,,\\qquad\\text{\\rm ran}(W_{\\pm}(A_{2},A_{1}))=\\mathsf H_{2}^{ac }\\,,\n$$\n$$ W_{\\pm}(A_ {1},A_{1})^{*}=W_{\\pm}(A_{2},A_ {1})\\,,\n$$ \nwhere $P_{k}^{ac}$ denotes the orthogonal projector onto the absolutely continuous subspace $\\mathsf H_{k}^{ac}$ of $A_{k}$. Furthermore, we say the asymptotic completeness holds for the scattering couple $(A_{1},A_{2})$ whenever, beside completeness, one has \n$$\n\\mathsf H_{1}^{ac }=(\\mathsf H_{1}^{pp})^{\\perp}\\,,\\qquad \\mathsf H_{2}^{ac }=(\\mathsf H_{2}^{pp})^{\\perp}\\,,\n$$ \nwhere $\\mathsf H_{k}^{pp}$ denotes the pure point subspace of $A_{k}$; equivalently, whenever $\n\\sigma_{sc}(A_{1})=\\sigma_{sc}(A_{2})=\\varnothing $.\n\\end{definition} \nOur next hypothesis is \\vskip8pt\\noindent\n(H6) completeness hold for the scattering couple $(A_{B_{1}},A)$. \n\\vskip8pt\\noindent\n\\begin{theorem}\\label{AC} Suppose that (H1)-(H6) hold. Then completeness holds for the couple $(A_{\\mathsf B},A)$. If furthermore ${e}(A_{B_{1}})$ is a discrete set and $\\sigma_{sc}(A)=\\varnothing $, then asymptotic completeness holds for the couple $(A_{\\mathsf B},A)$.\n\\end{theorem}\n\\begin{proof} By \\eqref{alt-res} and by the same proof as in Lemma \\ref{le-green}, one gets \n\\begin{equation}\\label{alt-gf}\n\\langle u,\\Delta_{\\mathsf B}v\\rangle_{L^{2}(M)}-\\langle A_{B_{1}}\nu,v\\rangle_{L^{2}(M)}=\n\\langle\\tau_{2} u,\\widehat \\rho_{\\mathsf B} v\\rangle_{\\mathfrak h_{2},\\mathfrak h_{2}^{*}}\\,,\\quad u\\in \\text{\\rm dom}(A_{B_{1}}),\\,v\\in\n\\text{\\rm dom}(A_{\\mathsf B})\\,,\n\\end{equation}\nwhere\n\\begin{equation*}\n\\widehat\\rho_{\\mathsf B}:\\text{\\rm dom}(A_{\\mathsf B})\\rightarrow\\mathfrak h_{2}^{*}\\,,\\quad\n\\widehat\\rho_{\\mathsf B}(R_{z}^{\\mathsf B}u):=\\widehat\\Lambda_{z}^{\\mathsf B}G^{B_{1}*}_{\\bar{z}}u\\,,\\qquad u\\in \\mathsf H\\,, \\quad z\\in\\varrho(A_{\\mathsf B})\\cap\\varrho(A_{B_{1}})\\,.\n\\end{equation*}\nThen, by hypotheses (H1)-(H5) and by \\cite[Theorems 2.8 and 3.8]{JMPA} (compare \\eqref{alt-gf} and Lemma \\ref{bound} here with (2.19) and Lemma 3.6 there and notice that hypothesis (H6) there is included in our hypothesis (H4)) one gets the completeness for the couple $(A_{\\mathsf B},A_{B_{1}})$. By (H6) and the chain rule for the wave operators (see \\cite[Theorem 3.4, Chapter X]{Kato}), one then gets completeness for the scattering couple $(A_{\\mathsf B},A)$. To conclude the proof it remains to show that \n$\\sigma_{sc}(A_{\\mathsf B})=\\varnothing$. Let $\\mathsf H_{\\mathsf B}^{pp}$ denote the pure point subspace of $A_{\\mathsf B}$; given $u\\in (\\mathsf H_{\\mathsf B}^{pp})^{\\perp}$, let $\\mu_{u}^{\\mathsf B}$ be the corresponding spectral measure. By Theorem \\ref{LAP} and by standard arguments (see, e.g., the proof of \\cite[Thm. 6.1]{Agmon}), the support of the singular continuous component of $\\mu_{u}^{\\mathsf B}$ is contained in $e(A_{B_{1}})$. Since $e(A_{B_{1}})$ is discrete, such a support is empty and so $u$ has a null projection onto $\\mathsf H_{\\mathsf B}^{sc}$, the singular continuous subspace of $A_{\\mathsf B}$. This gives $(\\mathsf H_{\\mathsf B}^{pp})^{\\perp}=\\mathsf H_{\\mathsf B}^{ac}$, where $\\mathsf H_{\\mathsf B}^{ac}$ denote the absolutely continuous subspace of $A_{\\mathsf B}$.\n\\end{proof}\n\n\\subsection{A representation formula for the scattering matrix}\nAccording to Theorem \\ref{AC}, under the assumptions there stated, the scattering operator \n$$\nS_{\\mathsf B}:=W_{+}(A_{\\mathsf B},A)^{*}W_{-}(A_{\\mathsf B},A)\n$$\nis a well defined unitary map. Let \n\\begin{equation}\\label{F0}\nF:L^{2}(M)_{ac}\\to\\int^{\\oplus}_{\\sigma_{ac}(A)}(L^{2}(M)_{ac})_{\\lambda}\\,d\\eta(\\lambda)\n\\end{equation}\nbe a unitary map which diagonalizes the absolutely continuous component of $A$, i.e., a direct integral representation of $L^{2}(M)_{ac}$, the absolutely continuous subspace relative to $A$, with respect to the spectral measure of the absolutely continuous component of $A$ (see e.g. \\cite[Section 4.5.1]{BW}). We define the scattering matrix $${\\mathcal S}^{\\mathsf B}_{\\lambda}:(L^{2}(M)_{ac})_{\\lambda}\\to (L^{2}(M)_{ac})_{\\lambda}$$ by the relation (see e.g. \\cite[Section 9.6.2]{BW})\n$$\nFS_{\\mathsf B}F^{*}u_{\\lambda}={\\mathcal S}^{\\mathsf B}_{\\lambda}u_{\\lambda}\n\\,.\n$$\nNow, following the same scheme as in \\cite{JMPA}, which uses the Birman-Kato invariance principle and the Birman-Yafaev general scheme in stationary scattering theory, we provide an explicit relation between ${\\mathcal S}^{\\mathsf B}_{\\lambda}$ and $\\Lambda^{\\!\\mathsf B,+}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\Lambda^{\\!\\mathsf B}_{\\lambda+i\\epsilon}$. \\par \nGiven $\\mu\\in \\varrho(A)\\cap\\varrho(A_{\\mathsf B})$, we consider the scattering couple $(R^{\\mathsf B}_{\\mu}, R_{\\mu})$ and the strong limits \n$$\nW_{\\pm}(R^{\\mathsf B}_{\\mu},R_{\\mu})\n:=\\text{s-}\\lim_{t\\to\\pm\\infty}e^{itR^{\\mathsf B}_{\\mu}}e^{-itR_{\\mu}}P^{\\mu}_{ac}\\,,\n$$ \nwhere $P^{\\mu}_{ac}$ is the orthogonal projector onto the absolutely continuous subspace of $R_{\\mu}$; we prove below that such limits exist everywhere in $L^{2}(M)$. Let $S^{\\mu}_{\\Lambda}$ the corresponding scattering operator \n$$\nS_{\\mathsf B}^{\\mu}:=W_{+}(R^{\\mathsf B}_{\\mu},R_{\\mu})^{*}W_{-}(R^{\\mathsf B}_{\\mu},R_{\\mu})\\,.\n$$\nUsing the unitary operator $F_{\\mu}$ which diagonalizes the absolutely continuous component of $R_{\\mu}$, i.e. $(F_{\\mu}u)_{\\lambda}:=\\frac1\\lambda(Fu)_{\\mu-\\frac1\\lambda}$, $\\lambda\\not=0$ such that $\\mu-\\frac1\\lambda\\in \\sigma_{ac}(A)$, one defines the scattering matrix $${\\mathcal S}^{\\mathsf B,\\mu}_{\\lambda}:(L^{2}(M)_{ac})_{\\mu-\\frac1\\lambda}\\to (L^{2}(M)_{ac})_{\\mu-\\frac1\\lambda}$$ corresponding to the scattering operator $S^{\\mu}_{\\mathsf B}$ by the relation\n$$\nF_{\\mu}S^{\\mu}_{\\mathsf B}F_{\\mu}^{*}u^{\\mu}_{\\lambda}={\\mathcal S}^{\\mathsf B,\\mu}_{\\lambda}u^{\\mu}_{\\lambda}\n\\,.\n$$ \nWe introduce a further hypothesis (H7), which we split in four separate points:\\vskip5pt\\par\\noindent\n(H7.1) $A$ is bounded from above and satisfies a Limiting Absorption Principle: there exists a (eventually empty) closed set ${e}(A)\\subset\\mathbb{R}$ of zero Lebesgue measure such that for all $\\lambda\\in \\mathbb{R}\\backslash{e}(A)$ the limits\n\\begin{equation}\\label{LAP0}\nR^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}R_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(L_{\\varphi}^{2}(M),L_{\\varphi^{-1}}^{2}(M))$; \n\\vskip5pt\\par\\noindent\n(H7.2) $G^{1}_{z}\\in\\mathscr B(\\mathfrak h_{1}^{*},L^{2}_{\\varphi}(M))$ for any $z\\in\\varrho(A)$ and the limits\n\\begin{equation}\\label{limG1}\nG_{\\lambda}^{{1},\\pm}:=\\lim_{\\epsilon\\searrow 0}G^{{1}}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(\\mathfrak h_{1} ^{*},L^{2}_{\\varphi^{-1}}(M))$ for any $\\lambda\\in\\mathbb{R}\\backslash{e}(A)$;\n\\vskip5pt\\par\\noindent\n(H7.3) the limits\n\\begin{equation}\\label{limLB1}\n\\Lambda^{\\!B_{1},\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\Lambda^{\\!B_{1},\\pm}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(\\mathfrak h_{1},\\mathfrak h_{1}^{*})$ for any $\\lambda\\in\\mathbb{R}\\backslash{e}(A_{B_{1}})$;\n\\vskip5pt\\par\\noindent\n(H7.4) the limits\n\\begin{equation}\\label{limtG1}\n\\tau_{2} G^{1,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\tau_{2} G^{1}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(\\mathfrak b^{*}_{1},\\mathfrak b_{2})$ for any $\\lambda\\in\\mathbb{R}\\backslash{e}(A_{B_{1}})$.\n\\begin{remark}\\label{rem3.6} By $\\tau_{2}G^{1}_{z}=\\tau_{2} (\\tau_{1}R_{\\bar z})^{*}=(\\tau_{1} (\\tau_{2}R_{ z})^{*})^{*}=(\\tau_{1}G^{2}_{\\bar z})^{*}$, hypothesis (H7.4) entails the existence in $\\mathscr B(\\mathfrak b_{2},\\mathfrak b_{1}^{*})$, for any $\\lambda\\in\\mathbb{R}\\backslash{e}(A_{B_{1}})$, of the limits \n\\mathfrak b\n\\tau_{1} G^{2,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\tau_{1} G^{2}_{\\lambda\\pm i\\epsilon}\\,.\n\\end{equation}\n\\end{remark}\n\\begin{remark} Whenever one strengthens hypotheses (H7.2) as in (H5), then, by the same kind of proof that leads to the existence of the limit \\eqref{limLambda} (see \\cite[Lemma 3.6]{JMPA}), one gets the existence of the limits requested in hypotheses (H7.3).\n\\end{remark}\n\\begin{lemma}\\label{rmH7} Suppose that (H1)-(H5) and (H7) hold. Then\n\\begin{equation}\\label{rmH7-1}\nR_{\\lambda}^{B_{1},\\pm}=R^{\\pm}_{\\lambda}+G_{\\lambda}^{{1},\\pm}\\Lambda^{\\!B_{1},\\pm}_{\\lambda}(G_{\\lambda}^{{1},\\mp})^{*}\\,;\n\\end{equation}\n\\begin{equation}\\label{Gz2}\nG^{2}_{z}\\in \\mathscr B(\\mathfrak h_{2} ^{*},L^{2}_{\\varphi}(M))\\,,\\qquad z\\in\\varrho(A_{B_{1}})\\cap\\varrho(A)\\,,\n\\end{equation} \nthe limits\n\\begin{equation}\\label{limG2}\nG_{\\lambda}^{{2},\\pm}:=\\lim_{\\epsilon\\searrow 0}G^{{2}}_{\\lambda\\pm i\\epsilon}\n\\end{equation}\nexist in $\\mathscr B(\\mathfrak h_{2} ^{*},L^{2}_{\\varphi^{-1}}(M))$ for any $\\lambda\\in\\mathbb{R}\\backslash{e}(A_{B_{1}})$ and\n\\begin{equation}\\label{limGB1}\nG_{\\lambda}^{B_{1},\\pm}=G_{\\lambda}^{2,\\pm}+G_{\\lambda}^{1,\\pm}\\Lambda^{\\!B_{1},\\pm}_{\\lambda}\\tau_{1}G_{\\lambda}^{2,\\pm}\\,;\n\\end{equation}\nthe limits \n$$\n\\Lambda_{\\lambda}^{\\!\\mathsf B,\\pm}:=\n\\lim_{\\epsilon\\searrow 0}\\Lambda_{\\lambda\\pm i\\epsilon}^{\\!\\mathsf B}\n$$\nexist in $\\mathscr B(\\mathfrak h_{1}\\oplus \\mathfrak b_{2},\\mathfrak h_{1}^{*}\\oplus \\mathfrak b_{2}^{*})$ and\n\\begin{align}\n\\Lambda_{\\lambda}^{\\!\\mathsf B,\\pm}=&\\begin{bmatrix}\\Lambda^{\\!B_{1},\\pm}_{\\lambda}+\n\\Lambda^{\\!B_{1},\\pm}_{\\lambda}\\tau_{1}G^{2,\\pm}_{\\lambda}\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\\tau_{2}G^{1,\\pm}_{\\lambda}\\Lambda^{\\!B_{1},\\pm}_{\\lambda}&\n\\Lambda^{\\!B_{1},\\pm}_{\\lambda}\\tau_{1}G^{2,\\pm}_{\\lambda}\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\\\\\n\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\\tau_{2}G^{1,\\pm}_{\\lambda}\\Lambda_{\\lambda}^{\\!B_{1},\\pm}&\n\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\n\\end{bmatrix}\\label{LBpm}\\\\\n=&\\left(1+\\begin{bmatrix}\\Lambda^{\\!B_{1},\\pm}_{\\lambda}&0\\\\0&\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\n\\end{bmatrix}\\begin{bmatrix}\\tau_{1}G^{2,\\pm}_{\\lambda}\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\\tau_{2}G^{1,\\pm}_{\\lambda}&\n\\tau_{1}G^{2,\\pm}_{\\lambda}\\\\\n\\tau_{2}G^{1,\\pm}_{\\lambda}&0\n\\end{bmatrix}\\,\\right)\\begin{bmatrix}\\Lambda^{\\!B_{1},\\pm}_{\\lambda}&0\\\\0&\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\n\\end{bmatrix}\\,.\\label{LBpm2}\n\\end{align}\n\\end{lemma}\n\\begin{proof} The relation \\eqref{rmH7-1} is an immediate consequence of \\eqref{res1.1} and (H7.1)-(H7.3). By \\eqref{GB1}, \n$$\nG_{z}^{2}=G^{B_{1}}_{z}-G_{z}^{1}\\Lambda^{\\!B_{1}}_{z}\\tau_{1}G_{z}^{2}\n$$\nand \\eqref{Gz2} follows from (H4.2) and (H7.2). Then, Remark \\ref{rem3.6}, (H.5) and (H7.3) entail \n\\eqref{limG2} and \\eqref{limGB1}. Finally, \\eqref{LBpm} and \\eqref{LBpm2} are consequence of \\eqref{LB-new}, \\eqref{LB-new2}, Corollary \\ref{limbb}, (H7.3), Remark \\ref{rem3.6} and (H7.4). \n\\end{proof}\nBefore stating the next results, let us notice the relations \n\\begin{equation}\\label{RR}\n\\left(-R_{\\mu}+z\\right)^{-1}=\\frac1z\\,\\left(1+\\frac1{z}\\,R_{\\mu-\\frac1z}\\right)\\,,\n\\quad \\left(-R_{\\mu}^{\\mathsf B}+z\\right)^{-1}=\\frac1z\\,\\left(1+\\frac1{z}\\,R^{\\mathsf B}_{\\mu-\\frac1z}\\right)\\,,\n\\end{equation}\nTherefore, by (H7.1) and Theorem \\ref{LAP}, the limits \n\\begin{equation}\\label{RRlim1}\n\\left(-R_{\\mu}+(\\lambda\\pm i0)\\right)^{-1}:=\\lim_{\\epsilon\\searrow 0}\\left(-R_{\\mu}+(\\lambda\\pm i\\epsilon)\\right)^{-1}\\,,\\quad \\lambda\\not=0\\,,\\ \\mu-\\frac1\\lambda\\in\\mathbb{R}\\backslash{e}(A)\\,,\n\\end{equation} \n\\begin{equation}\\label{RRlim2}\n\\left(-R_{\\mu}^{\\mathsf B}+(\\lambda\\pm i0)\\right)^{-1}:=\n\\lim_{\\epsilon\\searrow 0}\\left(-R_{\\mu}^{\\mathsf B}+(\\lambda\\pm i\\epsilon)\\right)^{-1}\\,,\n\\quad \\lambda\\not=0\\,,\\ \\mu-\\frac1\\lambda\\in\\mathbb{R}\\backslash{e}(A_{\\mathsf B})\\,,\n\\end{equation}\nexist in $\\mathscr B(L^{2}_{\\varphi}(M),L^{2}_{\\varphi^{-1}}(M))$.\n\\begin{theorem}\\label{BK}\nSuppose that hypotheses (H1)-(H7) hold. Then the strong limits \n\\begin{equation}\\label{WR}\nW_{\\pm}(R^{\\mathsf B}_{\\mu},R_{\\mu})\n:=\\text{s-}\\lim_{t\\to\\pm\\infty}e^{itR^{\\mathsf B}_{\\mu}}e^{-itR_{\\mu}}P^{\\mu}_{ac}\n\\end{equation}\nexist everywhere in $L^{2}(M)$.\nMoreover, for any $\\lambda\\not=0$ such that $\\mu-\\frac1\\lambda\\in \\sigma_{ac}(A)\\cap(\\mathbb{R}\\backslash{e}(A_{\\mathsf B}))$, one has\n\\begin{equation}\\label{S1}\n{\\mathcal S}^{\\mathsf B,\\mu}_{\\lambda}=1-2\\pi i\\,\\mathcal L^{\\mu}_{\\lambda}\n\\Lambda^{\\!\\mathsf B}_{\\mu}\\big(1+G^{*}_{\\mu}\\big(-R_{\\mu}^{\\mathsf B}+(\\lambda+ i0)\\big)^{-1}G_{\\mu}\\Lambda^{\\!\\mathsf B}_{\\mu}\\big)\n(\\mathcal L^{\\mu}_{\\lambda})^{*}\\,,\n\\end{equation}\nwhere\n\\begin{equation}\\label{SR}\n\\mathcal L^{\\mu}_\\lambda: \\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*}\\to(L^{2}(M)_{ac})_{\\mu-\\frac1\\lambda}\\,,\\quad \n\\mathcal L^{\\mu}_{\\lambda}(\\phi_{1}\\oplus\\phi_{2}):=\\frac1\\lambda(FG_{\\mu}(\\phi_{1}\\oplus\\phi_{2}))_{\\mu-\\frac1\\lambda}\\,.\n\\end{equation}\n\\end{theorem} \n\\begin{proof}\nBy \\eqref{resolvent}, one has $R_{\\mu}^{\\mathsf B}-R_{\\mu}=G_{\\mu}\\Lambda^{\\! B}_{\\mu}G^{*}_{\\mu}$ and we can use \\cite[Theorem 4', page 178]{Y} (notice that the maps there denoted by $G$ and $V$ corresponds to our $G_{\\mu}^{*}$ and $\\Lambda^{\\! B}_{\\mu}$ respectively). Let us check that the hypotheses there required are satisfied. Since $G^{*}_{\\mu}\\in \\mathscr B(L^{2}(M),\\mathfrak h_{1}\\oplus\\mathfrak h_{2})$, the operator $G_{\\mu}$ is $|R_{\\mu}|^{1\/2}$-bounded. By (H7.2) and \\eqref{Gz2}, one has $G_{z}\\in\\mathscr B(\\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*}, L^{2}_{\\varphi}(M))$ for any $z\\in\\varrho(A_{B_{1}})\\cap\\varrho(A)\\supset[\\lambda_{1},+\\infty)\\ni\\mu$. Therefore, by \\eqref{RRlim1}, \\eqref{RRlim2}, (H7.1), Theorem \\ref{LAP} and (H4), the limits \n$$\n\\lim_{\\epsilon\\searrow 0}\\, G^{*}_{\\mu}(-R _{\\mu}+(\\lambda\\pm i\\epsilon))^{-1}\\,,\n$$\n$$\n\\lim_{\\epsilon\\searrow 0}\\, G^{*}_{\\mu}(-R^{\\mathsf B}_{\\mu}+(\\lambda\\pm i\\epsilon))^{-1}\\,,\n$$\n$$\n\\lim_{\\epsilon\\searrow 0}\\, G^{*}_{\\mu}(-R^{\\mathsf B}_{\\mu}+(\\lambda\\pm i\\epsilon))^{-1}G_{\\mu}\n$$\nexist. Therefore, to get the thesis we need to check the validity of the remaining hypothesis in \n\\cite[Theorem 4', page 178]{Y}: $G^{*}_{\\mu}$ is weakly-$R _{\\mu}$ smooth, i.e., by \\cite[Lemma 2, page 154]{Y}, \n\\begin{equation}\\label{in1.1}\n\\sup_{0<\\epsilon<1}\\epsilon\\,{\\|}G^{*}_{\\mu} (-R _{\\mu}+(\\lambda\\pm i\\epsilon))^{-1}{\\|}_{L^{2}(M),\\mathfrak h_{1}\\oplus\\mathfrak h_{2}}^{2}\\le c_{\\lambda}<+\\infty\\,,\\quad\\text{a.e. $\\lambda$}\\,.\n\\end{equation}\nBy \\eqref{RR}, this is consequence of\n\\begin{equation}\\label{in2.2}\n\\sup_{0<\\epsilon<1}\\epsilon\\,{\\|}G^{*}_{\\mu}R_{\\mu-\\frac1\\lambda\\pm i\\epsilon}{\\|}_{L^{2}(M),\\mathfrak h_{1}\\oplus\\mathfrak h_{2}}^{2}\\le C_{\\lambda}<+\\infty\\,,\\quad\\text{a.e. $\\lambda$}\\,.\n\\end{equation}\nBy \\cite[(3.16)]{JMPA},\n\\begin{align*}\n&\\epsilon\\,{\\|}G_{\\lambda\\pm i\\epsilon}{\\|}_{\\mathfrak h^{*}_{1}\\oplus\\mathfrak h^{*}_{2},L^{2}(M)}^{2}\\\\\n\\le& \n\\frac12\\,(|\\mu-\\lambda|+\\epsilon)\\, {\\|}G_{\\mu}{\\|}_{\\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*},L_{\\varphi}^{2}(M)} \\left(\\|G_{\\lambda- i\\epsilon}\\|_{\\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*},L_{\\varphi^{-1}}^{2}(M)} +\n\\|G_{\\lambda+ i\\epsilon}\\|_{\\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*},L_{\\varphi^{-1}}^{2}(M)} \\right)\\,.\n\\end{align*}\nThen, \\eqref{in2.2} follows from \\eqref{limG1}, \\eqref{limG2} and the equality\n\\begin{align*}\n&{\\|}G^{*}_{\\mu} R _{z}{\\|}_{L^{2}(M),\\mathfrak h_{1}\\oplus\\mathfrak h_{2}}\n={\\|}\\tau R_{\\mu}R _{z}{\\|}_{L^{2}(M),\\mathfrak h_{1}\\oplus\\mathfrak h_{2}}= \n{\\|}\\tau R _{z}R _{\\mu}{\\|}_{L^{2}(M),\\mathfrak h_{1}\\oplus\\mathfrak h_{2}}\\\\\n=&\n{\\|}R _{\\mu}(\\tau R _{z})^{*}{\\|}_{\\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*},L^{2}(M)}\\le\n{\\|}R _{\\mu}{\\|}_{L^{2}(M),L^{2}(M)} {\\|}G_{\\bar z}{\\|}_{\\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*},L^{2}(M)}\\,.\n\\end{align*}\nThus, by \\cite[Theorem 4', page 178]{Y}, \nthe limits \\eqref{WR} exist everywhere in $L^{2}(M)$ and the corresponding scattering matrix is given by \\eqref{S1}, where $\\mathcal L^{\\mu}_{\\lambda}\\phi:=(F^{\\mu}G_{\\mu}\\phi)_{\\lambda}=\\frac1\\lambda(FG_{\\mu}\\phi)_{\\mu-\\frac1\\lambda}$. \n\\end{proof}\n\\begin{theorem}\\label{S-matrix} Suppose that hypotheses (H1)-(H7) hold. Then the scattering matrix of the couple $(A_{\\mathsf B},A)$ has the representation\n\\begin{equation}\\label{S-M}\n{\\mathcal S}^{\\mathsf B}_{\\lambda}=1-2\\pi i\\mathcal L_{\\lambda}\\Lambda^{\\!\\mathsf B,+}_{\\lambda}\\mathcal L_{\\lambda}^{*}\\,,\\quad \\lambda\\in\\sigma_{ac}(A)\\cap(\\mathbb{R}\\backslash{e}(A_{\\mathsf B}))\\,,\n\\end{equation}\nwhere $\\mathcal L_\\lambda: \\mathfrak h_{1}^{*}\\oplus\\mathfrak h_{2}^{*}\\to(L^{2}(M)_{ac})_{\\lambda}$ is the $\\mu$-independent linear operator defined by \n$$\n\\mathcal L_{\\lambda}(\\phi_{1}\\oplus\\phi_{2}):=(\\mu-\\lambda)(FG_{\\mu}(\\phi_{1}\\oplus\\phi_{2}))_{\\lambda}\n$$\nand $\\Lambda^{\\!\\mathsf B,+}_{\\lambda}$ is given in \\eqref{LBpm}.\n\\end{theorem}\n\\begin{proof}\nBy Theorem \\ref {AC}, Theorem \\ref{BK} and by Birman-Kato invariance principle (see e.g. \\cite[Section II.3.3]{BW}), one has\n$$\nW_{\\pm}(A_{\\mathsf B},A)=W_{\\pm}(R^{\\mathsf B}_{\\mu},R_{\\mu})\n$$\nand so\n$$\nS_{\\mathsf B}=S_{\\mathsf B}^{\\mu}\\,.\n$$\nThus, since $(F^{\\mu}u)_{\\lambda}=\\frac1\\lambda(Fu)_{\\mu-\\frac1\\lambda}$, one obtains (see also \\cite[Equation (14), Section 6, Chapter 2]{Y})\n\\begin{equation}\\label{SS}\n{\\mathcal S}^{\\mathsf B}_{\\lambda}={\\mathcal S}^{\\mathsf B,\\mu}_{(-\\lambda+\\mu)^{-1}}\\,.\n\\end{equation}\nBy \\cite[Lemma 4.2]{JMPA}, for any $z\\not=0$ such that $\\mu-\\frac1z\\in \\varrho(A_{\\mathsf B})\\cap\\varrho(A)$, there holds\n$$\n\\Lambda^{\\!\\mathsf B}_{\\mu}\\left(1+G^{*}_{\\mu}\\left(-R_{\\mu}^{\\mathsf B}+z\\right)^{-1}G_{\\mu}\\Lambda^{\\! B}_{\\mu}\\right)=\\Lambda^{\\!\\mathsf B}_{\\mu-\\frac1z}\\,.\n$$\nHence, whenever $z=\\lambda\\pm i\\epsilon$ and $\\mu-\\frac1{\\lambda}\\in\\mathbb{R}\\backslash{e}(A_{\\mathsf B})$, one gets, as $\\epsilon\\downarrow 0$, \n$$\n\\Lambda^{\\!\\mathsf B}_{\\mu}\\big(1+G^{*}_{\\mu}\\left(-R_{\\mu}^{\\mathsf B}+(\\lambda\\pm i0)\\right)^{-1}G_{\\mu}\\Lambda^{\\!\\mathsf B}_{\\mu}\\big)=\\Lambda^{\\!\\mathsf B,\\pm}_{\\mu-\\frac1{\\lambda}}\\,.\n$$\nThe proof is then concluded by Theorem \\ref{BK}, by \\eqref{SS} and by setting \n$\\mathcal L_{\\lambda}:=\\mathcal L^{\\mu}_{{(-\\lambda+\\mu)^{-1}}}$. The operator $\\mathcal L_{\\lambda}$ is $\\mu$-independent by invariance principle (see the proof in \\cite[Corollary 4.3]{JMPA} for an explicit check).\n\\end{proof}\n\\begin{remark}\nBy \\eqref{LBpm}, \n$$\n\\Lambda^{\\!\\mathsf B,\\pm}_{\\lambda}=\\begin{bmatrix}\\Lambda^{\\!B_{1},\\pm}_{z}&0\\\\0&0\\end{bmatrix}+\\widetilde \\Lambda^{\\!\\mathsf B,\\pm}_{\\lambda}\\,,\n$$\nwhere\n$$\n\\widetilde \\Lambda^{\\!\\mathsf B,\\pm}_{\\lambda}:=\n\\begin{bmatrix}\n\\Lambda^{\\!B_{1},\\pm}_{\\lambda}\\tau_{1}G^{2,\\pm}_{\\lambda}\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\\tau_{2}G^{1,\\pm}_{\\lambda}\\Lambda^{\\!B_{1},\\pm}_{\\lambda}&\n\\Lambda^{\\!B_{1},\\pm}_{\\lambda}\\tau_{1}G^{2,\\pm}_{\\lambda}\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\\\\\n\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\\tau_{2}G^{1,\\pm}_{\\lambda}\\Lambda_{\\lambda}^{\\!B_{1},\\pm}&\n\\widehat \\Lambda^{\\mathsf B,\\pm}_{\\lambda}\n\\end{bmatrix}\\,.\n$$\nTherefore, defining \n$$\n\\mathcal L^{1}_{\\lambda}\\phi_{1}:=\\mathcal L_{\\lambda}(\\phi_{1}\\oplus 0)\\,,\n$$ \none gets\n$$\n{\\mathcal S}^{\\mathsf B}_{\\lambda}={\\mathcal S}^{B_{1}}_{\\lambda}-2\\pi i\\mathcal L_{\\lambda}\\widetilde\\Lambda^{\\!\\mathsf B,+}_{\\lambda}\\mathcal L_{\\lambda}^{*}\\,,\n$$\nwhere\n\\begin{equation}\\label{SB1}\n{\\mathcal S}^{B_{1}}_{\\lambda}=1-2\\pi i\\mathcal L^{1}_{\\lambda}\\widetilde\\Lambda^{\\!B_{1},+}_{\\lambda}(\\mathcal L^{1}_{\\lambda})^{*}\n\\end{equation}\nis the scattering matrix relative to the couple $(A_{B_{1}},A)$. Moreover, in the case $B_{1}=0$, \ndefining \n$$\n\\mathcal L^{2}_{\\lambda}\\phi_{2}:=\\mathcal L_{\\lambda}(0\\oplus\\phi_{2})\\,,\n$$ \none gets the following representation formula for the scattering couple $(A_{B_{0},B_{2}},A)$ (compare with \\cite[Corollary 4.3]{JMPA}):\n$$\n{\\mathcal S}^{B_{0},B_{2}}_{\\lambda}=1-2\\pi i\\mathcal L^{2}_{\\lambda}\\Lambda^{\\!B_{0},B_{2},+}_{\\lambda}(\\mathcal L^{2}_{\\lambda})^{*}\\,.\n$$\nLet us further notice that, whenever $A$ is the free Laplacian in $L^{2}(\\mathbb{R}^{3})$ and $B_{1}$ corresponds to a perturbation by a regular potential as in Section 5 below, then \\eqref{SB1} gives the usual formula for the scattering matrix for a short-range potential (see, e.g., \\cite[Section 8]{Y-LNM}).\n\\end{remark}\n\n\\section{Kato-Rellich perturbations and their layers potentials}\n\n\\subsection{\\label{Sec_V}Potential perturbations}\n\nIn this section we suppose that the real-valued potential $\\mathsf v$ is of Kato-Rellich type, i.e., $\\mathsf v\\in L^{2}(\\mathbb{R}^{3})+L^{\\infty}(\\mathbb{R}^{3})$, equivalently,\n\\begin{equation}\n\\mathsf v=\\mathsf v_{2}+\\mathsf v_{\\infty} \\,,\\qquad \\mathsf v_{2}\\in L^{2}( \\mathbb{R}^{3}) \\,,\\qquad \\mathsf v_{\\infty}\\in L^{\\infty}\n(\\mathbb{R}^{3}) \\,. \\label{K-R}%\n\\end{equation}\nWe use the same simbol $\\mathsf v$ to denote both the potential function and the corresponding multiplication operator $u\\mapsto \\mathsf v u$. \n\\par\nGiven $\\Omega\\subset\\mathbb{R}^{3}$, open and bounded with a Lipschitz boundary $\\Gamma$, we define $H^{s}(\\mathbb{R}^{3}\\backslash\\Gamma)\\hookleftarrow H^{s}(\\mathbb{R}^{3})$ by \n$$\nH^{s}(\\mathbb{R}^{3}\\backslash\\Gamma):=H^{s}(\\Omega)\\oplus H^{s}(\\Omega_{\\rm ex})\\,,\\qquad s\\ge 0\\,.\n$$\nWe refer to \\cite[Chapter 3]{McL} for the definition of the Sobolev spaces $H^{s}(\\mathbb{R}^{3})$, $H^{s}(\\Omega)$ and $H^{s}(\\Gamma)$. One has \n$$\nH^{s}(\\mathbb{R}^{3}\\backslash\\Gamma)=H^{s}(\\mathbb{R}^{3})\\,,\\qquad 0\\le s<1\/2\\,.\n$$\nSince (see \\cite[Theorems 3.29 and 3.30]{McL}),\n$$\nH^{s}( \\mathcal O) ^{\\ast}\n=H_{\\overline{\\mathcal O}}^{-s}( \\mathbb{R}^{3}) \\,,\\qquad\ns\\in\\mathbb{R}\\,,\n$$\n$H_{\\overline{\\mathcal O}}^{-s}( \\mathbb{R}^{3})$ denoting the set of distributions $H^{-s}( \\mathbb{R}^{3})$ with support in ${\\overline{\\mathcal O}}$,\none has\n\\begin{equation}\nH^{s}( \\mathbb{R}^{3}\\backslash\\Gamma)^{\\ast}=H^{s}( \\Omega)^{\\ast}\n\\oplus H^{s}( \\mathbb{R}^{3}\\backslash\\overline\\Omega)^{\\ast}=\nH^{-s}_{\\overline\\Omega}(\\mathbb{R}^{3})\\oplus H^{-s}_{\\Omega^{c}}(\\mathbb{R}^{3})\\hookrightarrow H^{-s}(\\mathbb{R}^{3})\n\\,.\n\\label{dual}%\n\\end{equation}\nLet us notice that\n\\begin{equation}\\label{B}\n\\mathscr B(H^{s}( \\mathbb{R}^{3}\\backslash\\Gamma),H^{t}( \\mathbb{R}^{3}\\backslash\\Gamma)^{\\ast})\n\\hookrightarrow\n\\mathscr B(H^{s}( \\mathbb{R}^{3}),H^{-t}( \\mathbb{R}^{3}))\\,,\\qquad s,t\\ge 0\\,,\n\\end{equation}\nand\n\\begin{equation}\\label{B*}\n\\mathscr B(H^{-s}( \\mathbb{R}^{3}),H^{t}( \\mathbb{R}^{3}))\\hookrightarrow\\mathscr B(H^{s}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*},H^{t}( \\mathbb{R}^{3}\\backslash\\Gamma))\n\\,,\\qquad s,t\\ge 0\\,.\n\\end{equation}\n\\begin{lemma}\\label{v}\n\\begin{equation}\\label{v-sob}\n\\mathsf v\\in{\\mathscr B}( H^{1+s}( \\mathbb{R}^{3}\\backslash\\Gamma) ,H^{1-s}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,,\\qquad -1\\le s\\le 1 \\,. \n\\end{equation}\n\\end{lemma}\n\\begin{proof}\nGiven $u= u_{\\rm in}\\oplus u_{\\rm ex}\\in H^{2}( \\mathbb{R}^{3}\\backslash\\Gamma) \n$ one has%\n\\[\n\\|\\mathsf v_{\\infty}u\\| _{L^{2}(\\mathbb{R}^{3})}\\leq\\|\\mathsf v\\|_{L^{\\infty}(\\mathbb{R}^{3})}\\| u\\| _{L^{2}(\\mathbb{R}^{3})}\\leq\\|\\mathsf v\\|_{L^{\\infty}(\\mathbb{R}^{3})}\\| u\\|_{H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma) }\\,.\n\\]\nand\n\\begin{align*}\n\\|\\mathsf v_{2}u\\|_{L^{2}(\\mathbb{R}^{3})}=&\\|\\mathsf v_{2}\\|_{L^{2}(\\Omega)}\\|u_{\\rm in}\\|_{L^{\\infty}(\\Omega)}+\\|\\mathsf v_{2}\\|_{L^{2}(\\mathbb{R}^{3}\\backslash\\overline\\Omega)}\n\\|u_{\\rm ex}\\|_{L^{\\infty}(\\mathbb{R}^{3}\\backslash\\overline\\Omega)}\\\\\n\\lesssim &\n\\|\\mathsf v_{2}\\|_{L^{2}(\\Omega)}\\|u_{\\rm in}\\|_{H^{2}(\\Omega)}+\n\\|\\mathsf v_{2}\\|_{L^{2}(\\mathbb{R}^{3}\\backslash\\overline\\Omega)}\n\\|u_{\\rm ex}\\|_{H^{2}(\\mathbb{R}^{3}\\backslash\\overline\\Omega)}\\\\\n\\lesssim &\\|\\mathsf v_{2}\\|_{L^{2}(\\mathbb{R}^{3})}\n\\|u\\|_{H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma)}\\,.\n\\end{align*}\nHence $\\mathsf v\\in{\\mathscr B}( H^{2}( \\mathbb{R}^{3}\\backslash\\Gamma) ,L^{2}( \\mathbb{R}^{3}))$. Then, for any $u,v\\in H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma)$, one has\n\\begin{align*}\n&\\big|\\langle \\mathsf v u,v\\rangle_{H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*},H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma)}\\big|\n=\\big|\\langle \\mathsf v u,v\\rangle_{L^{2}(\\mathbb{R}^{3})}\\big|\\\\\n=&\n\\big|\\langle u,\\mathsf v v\\rangle_{L^{2}(\\mathbb{R}^{3})}\\big|\n\\le \\|\\mathsf v\\|_{H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma),L^{2}(\\mathbb{R}^{3})}\\|u\\|_{L^{2}(\\mathbb{R}^{3})}\n\\|v\\|_{H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma)}\n\\end{align*}\nand so $u\\mapsto\\mathsf v u$ extends to a map in $\\mathscr B(L^{2}( \\mathbb{R}^{3}),H^{2}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*} )$. The proof is then concluded by interpolation.\n\\end{proof}\nIn the following, $R_{z}$ denotes the resolvent of the free Laplacian, i.e.,\n\\begin{equation}\\label{free}\nR_{z}:=\\left( -\\Delta+z\\right) ^{-1}\\in\\mathscr B(H^{s}(\\mathbb{R}^{3}),H^{s+2}(\\mathbb{R}^{3}))\\,,\\quad s\\in\\mathbb{R}\\,.\n\\end{equation}\nSince $\\mathsf v$ is of Rellich-Kato type, one has (see, e.g., \\cite[Section 3, $\\S$5, Chap. V]{Kato}): \n\\begin{theorem}\\label{KR}\n$\\Delta+\\mathsf v:H^{2}( \\mathbb{R}^{3})\\subset L^{2}( \\mathbb{R}^{3})\\to L^{2}( \\mathbb{R}^{3}) $ is\nself-adjoint and semi-bounded from above. Moreover, for $z\\in\\mathbb{C}$ sufficiently far away from $[0,+\\infty)$, $\\|\\mathsf v R_{z}\\|_{L^{2}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3})}<1$, \nand\n\\begin{equation}\nR_{z}^{\\mathsf v}:=(-(\\Delta+\\mathsf v)+z)^{-1}=R_{z}+R_{z}( 1- \\mathsf v R_{z}) ^{-1}\\mathsf v R_{z}\n\\,,\\label{Rvz}\n\\end{equation}\n\\begin{equation}\n( 1-\\mathsf v R_{z}) ^{-1}=\\sum_{k=0}^{+\\infty}\\left( \\mathsf v R_{z}\\right) ^{k}\n\\in{\\mathscr B}( L^{2}(\\mathbb{R}^{3}))\\,.\n\\label{L-v-est}%\n\\end{equation}\n\\end{theorem}\n\\begin{remark}\\label{tpt} Let us notice that the Kato-Rellich theorem could be obtained by Corollary \\ref{cor1} by taking $\\tau_{1}u:=u$ and $B_{1}=\\mathsf v$. Hence, \\eqref{Rvz} holds for any $z$ in $\\varrho(\\Delta+\\mathsf v)\\cap\\mathbb{C}\\backslash(-\\infty,0]$ and $( 1+\\mathsf v R_{z}) ^{-1}\\in {\\mathscr B}( L^{2}(\\mathbb{R}^{3}))$ there.\n\\end{remark}\n\\begin{remark}\\label{sa} \nBy \\eqref{free}, \\eqref{Rvz}, \\eqref{L-v-est}, \\eqref{v-sob} and \\eqref{B}, one has $R^{\\mathsf v}_{\\bar z}\\in\n\\mathscr B(L^{2}( \\mathbb{R}^{3}),H^{2}( \\mathbb{R}^{3}))$ and hence $(R_{\\bar z}^{\\mathsf v})^{*}\\in \\mathscr B(H^{-2}( \\mathbb{R}^{3}),L^{2}( \\mathbb{R}^{3}))$. Since $(\\Delta+\\mathsf v)$ is self-adjoint in $L^{2}(\\mathbb{R}^{3})$, $(R_{\\bar z}^{\\mathsf v})^{*}|L^{2}(\\mathbb{R}^{3})=\nR^{\\mathsf v}_{z}$. Therefore, $R^{\\mathsf v}_{z}:L^{2}(\\mathbb{R}^{3})\\subset H^{-2}(\\mathbb{R}^{3})\\to L^{2}(\\mathbb{R}^{3})$ extends to an operator in $\\mathscr B(H^{-2}( \\mathbb{R}^{3}),L^{2}( \\mathbb{R}^{3}))$ which, by abuse of notation, we still denote by $R_{ z}^{\\mathsf v}$ and which coincides with $(R_{\\bar z}^{\\mathsf v})^{*}$. Then, by interpolation, one gets\n\\begin{equation}\\label{Rvz-int}\nR_{z}^{\\mathsf v}\\in \\mathscr B(H^{s-1}( \\mathbb{R}^{3}),H^{s+1}( \\mathbb{R}^{3}))\\,,\\qquad -1\\le s\\le 1\\,.\n\\end{equation}\n\\end{remark}\n\\begin{remark} By \\eqref{Rvz}, \n\\begin{equation}\\label{btr}\n( 1- \\mathsf v R_{z}) ^{-1}\\mathsf v=(-\\Delta+z)R^{\\mathsf v}_{z}(-\\Delta+z)-(-\\Delta+z)\\,.\n\\end{equation}\nHence, by \\eqref{Rvz-int}, $( 1- \\mathsf v R_{z}) ^{-1}\\mathsf v\\in \\mathscr B(H^{2}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3}))$ extends to a map \n\\begin{equation}\\label{L-int}\n\\Lambda^{\\!\\mathsf v}_{z}\\in \\mathscr B(H^{s+1}( \\mathbb{R}^{3}),H^{s-1}( \\mathbb{R}^{3}))\\,,\n\\qquad -1\\le s\\le 1 \n\\end{equation}\nWith such a notation, $R_{z}^{\\mathsf v}$ in \\eqref{Rvz-int} has the representation\n\\begin{equation}\\label{RF-int}\nR_{z}^{\\mathsf v}=R_{z}+R_{z}\\Lambda^{\\!\\mathsf v}_{z}R_{z}\\,,\\qquad \n\\Lambda^{\\!\\mathsf v}_{z}|H^{2}(\\mathbb{R}^{3})=( 1- \\mathsf v R_{z}) ^{-1}\\mathsf v\\,.\n\\end{equation}\n\\end{remark}\n\\begin{remark}\\label{Lt} Since $\\|R_{z}\\mathsf v\\|_{L^{2}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3})}=\\|(R_{z}\\mathsf v)^{*}\\|_{L^{2}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3})}=\\|\\mathsf v R_{\\bar z}\\|_{L^{2}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3})}\n<1$ whenever $z\\in\\mathbb{C}$ is sufficiently far away from $[0,+\\infty)$, one has\n\\begin{equation}\\label{Lt2}\n( 1-R_{z}\\mathsf v) ^{-1}=\\sum_{k=0}^{+\\infty}\\left( R_{z}\\mathsf v\\right) ^{k}\n\\in{\\mathscr B}( L^{2}(\\mathbb{R}^{3}))\n\\end{equation}\nand\n\\begin{equation}\\label{Lt-2}\n\\mathsf v( 1- R_{z}\\mathsf v) ^{-1}\\in\\mathscr B(L^{2}(\\mathbb{R}^{3}),H^{-2}(\\mathbb{R}^{3}))\\,.\n\\end{equation}\nThen, \n$$\n\\big(( 1- R_{z}\\mathsf v) ^{-1}\\mathsf v\\big)^{*}=\n\\big(\\mathsf v( 1- R_{z}\\mathsf v)^{-1}\\big)^{*}=\\mathsf v(( 1- R_{z}\\mathsf v)^{*})^{-1}=\\mathsf v( 1- R_{\\bar z}\\mathsf v) ^{-1\n$$\nand so \n$$\n\\mathscr B(H^{-2}( \\mathbb{R}^{3}),L^{2}( \\mathbb{R}^{3}))\\ni(R^{\\mathsf v}_{z})^{*}=R_{\\bar z}+R_{\\bar z}\\mathsf v( 1- R_{\\bar z}\\mathsf v)^{-1}R_{\\bar z}=\nR^{\\mathsf v}_{\\bar z}=R_{\\bar z}+R_{\\bar z}\\Lambda_{\\bar z}^{\\!\\mathsf v}R_{\\bar z}\\,.\n$$\nTherefore \n\\begin{equation}\\label{LtL2}\n\\Lambda_{z}^{\\!\\mathsf v}|L^{2}(\\mathbb{R}^{3})=\\mathsf v( 1- R_{z}\\mathsf v) ^{-1}\\,.\n\\end{equation}\n\\end{remark}\n\\begin{lemma}\n\\begin{equation}\\label{Lvz-int}\n\\Lambda_{z}^{\\!\\mathsf v}\\in{\\mathscr B}( H^{1+s}(\\mathbb{R}^{3}\\backslash\\Gamma),\nH^{1-s}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*})\n\\,, \\qquad -1\\le s\\le 1\\,.\n\\end{equation}\n\\end{lemma}\n\\begin{proof} By Lemma \\ref{v} and by \\eqref{L-v-est}, one has $\\Lambda_{z}^{\\!\\mathsf v}=( 1+\\mathsf v R_{z}) ^{-1}\\mathsf v\\in{\\mathscr B}( H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma),L^{2}(\\mathbb{R}^{3}))$. By Lemma \\ref{v}, \\eqref{Lt2} and \\eqref{LtL2},\n$\\Lambda_{z}^{\\!\\mathsf v}\\in{\\mathscr B}( L^{2}(\\mathbb{R}^{3}), H^{2}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*})$.\nThe proof is then concluded by interpolation.\n\\end{proof}\nBy $H^{1-s}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\hookrightarrow H^{s-1}(\\mathbb{R}^{3})$ and \\eqref{free} one has\n\\begin{corollary}\n\\begin{equation}\\label{RLvz}\nR_{z}\\Lambda_{z}^{\\!\\mathsf v}\\in{\\mathscr B}( H^{s}(\\mathbb{R}^{3}\\backslash\\Gamma),\nH^{s}(\\mathbb{R}^{3}))\n\\,, \\qquad 0\\le s\\le 2\\,.\n\\end{equation}\n\n\\end{corollary}\nIn later proofs we will need the estimate provided in the following:\n\\begin{lemma}\nThere exist $c_{1}>0$, $c_{2}>0$ such that, for any $u\\equiv u_{\\rm in}\\oplus u_{\\rm ex}\\in H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)$ and for any $\\varepsilon>0$,\nthere holds\n\\begin{equation}\n\\big|\\langle \\mathsf v u,u\\rangle _{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*},H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)}\\big| \\leq\nc_{1}\\epsilon\\left(\\|\\nabla u_{\\rm in}\\|^{2}_{L^{2}(\\Omega_{\\rm in})}+\\|\\nabla u_{\\rm ex}\\|^{2}_{L^{2}(\\Omega_{\\rm ex})}\\right)+c_{2}(1+\\epsilon^{-3})\\|u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\,. \\label{V_est}%\n\\end{equation}\n\\end{lemma}\n\\begin{proof} By $H^{1}(\\Omega_{\\rm in\/\\rm ex}) \\hookrightarrow H^{3\/4}(\\Omega_{\\rm in\/\\rm ex}) \\hookrightarrow L^{4}(\\Omega_{\\rm in\/\\rm ex})$, by the Gagliardo-Niremberg \ninequalities (see \\cite{BM} for the interior case and \\cite{CM} for the exterior one)\n$$\n\\|u_{\\rm in}\\|_{L^{4}(\\Omega_{\\rm in})}\\lesssim\\|u_{\\rm in}\\|_{H^{3\/4}(\\Omega_{\\rm in})}\\lesssim\\|u_{\\rm in}\\|^{3\/4}_{H^{1}(\\Omega_{\\rm in})}\\|u_{\\rm in}\\|^{1\/4}_{L^{2}(\\Omega_{\\rm in})}\\,,\n$$ \n$$\n\\|u_{\\rm ex}\\|_{L^{4}(\\Omega_{\\rm ex})}\\lesssim\\|\\nabla u_{\\rm ex}\\|^{3\/4}_{L^{2}(\\Omega_{\\rm ex})}\\|u_{\\rm ex}\\|^{1\/4}_{L^{2}(\\Omega_{\\rm ex})}\n$$ \nand by Young's inequality\n$$\nxy\\le \\frac1{\\alpha}\\left({\\epsilon}\\ x^{\\alpha}+{(\\alpha-1)}\\,{\\epsilon^{-1\/(\\alpha-1)}}\n\\ y^{\\frac{\\alpha-1}{\\alpha}}\\right)\\,,\\qquad x,y,\\epsilon>0,\\, \\alpha>1\\,,\n$$\none gets\n$$\n\\|u\\|_{L^{4}(\\mathbb{R}^{3})}^{2}\\lesssim\n\\epsilon\\left(\\|\\nabla u_{\\rm in}\\|^{2}_{L^{2}(\\Omega_{\\rm in})}+\\|u\\|^{2}_{L^{2}(\\Omega_{\\rm in})}+\\|\\nabla u_{\\rm ex}\\|^{2}_{L^{2}(\\Omega_{\\rm ex})}\\right)+\\frac{1}{3}\\,\\epsilon^{-3}\\|u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\,.\n$$\nThe proof is then concluded by\n$$\n\\big|\\langle \\mathsf v u,u\\rangle _{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*},H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)}\\big| \\leq\n\\|\\mathsf v_{2}\\|_{L^{2}(\\mathbb{R}^{3})}\\|u\\|^{2}_{L^{4}(\\mathbb{R}^{3})}+\n\\|\\mathsf v_{\\infty}\\|_{L^{\\infty}(\\mathbb{R}^{3})}\\|u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\,.\n$$\n\\end{proof}\n\\begin{lemma}\\label{vH-1} For any $z\\in\\mathbb{C}$ sufficiently far away from $[0,+\\infty)$, one has \n$$\\|\\mathsf v R_{z}\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\!,H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}<1$$ and so\n\\begin{equation}\n( 1-\\mathsf v R_{z}) ^{-1}=\\sum_{k=0}^{+\\infty}\\left( \\mathsf v R_{z}\\right) ^{k}\n\\in{\\mathscr B}( H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,.\n\\end{equation}\n\\end{lemma}\n\\begin{proof} By \\eqref{V_est} and by the polarization identity, for any $u\\in H^{2}(\\mathbb{R}^{3})$ and $v$ in $H^{1}(\\mathbb{R}^{3})$ one has\n\\begin{align*}\n\\big|\\langle \\mathsf v u,v\\rangle_{L^{2}(\\mathbb{R}^{3})}\\big|\n \\leq&\n\\frac14\\Big(c_{1}\\epsilon\\,\\big|\\langle \\nabla u,\\nabla v\\rangle _{L^{2}(\\mathbb{R}^{3})}\\big|\n+c_{2}(1+\\epsilon^{-3})\\big|\\langle u,v\\rangle _{L^{2}(\\mathbb{R}^{3})}\\big|\\Big) \\\\\n =&\n\\frac14\\Big(c_{1}\\epsilon\\,\\big|\\langle -\\Delta u,v\\rangle _{L^{2}(\\mathbb{R}^{3})}\\big|\n+c_{2}(1+\\epsilon^{-3})\\big|\\langle u,v\\rangle _{L^{2}(\\mathbb{R}^{3})}\\big|\\Big)\\,. \n\\end{align*}\nBy the density of $L^{2}(\\mathbb{R}^{3})$ in $H^{1}(\\mathbb{R}^{3})$, the above inequality (here we refer to the second line) holds for any $v\\in L^{2}(\\mathbb{R}^{3})$. Then, by considering the supremum over the set of functions in $H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)$ with unitary norm, one gets\n\\begin{align*}\n\\|\\mathsf v u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}\n\\le&\\frac14\\Big( c_{1}\\epsilon\\,\\|-\\Delta u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}+\n\\big(c_{1}\\epsilon\\,|z|+c_{2}(1+\\epsilon^{-3})\\big)\\|u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}\\Big)\\\\\n\\le&\\frac14\\Big( c_{1}\\epsilon\\,\\|(-\\Delta+z) u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}+\n\\big(c_{1}\\epsilon\\,|z|+c_{2}(1+\\epsilon^{-3})\\big)\\|u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}\\Big)\\,.\n\\end{align*}\nTherefore\n$$\n\\|\\mathsf v R_{z}u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}\\le\n\\frac14\\Big( c_{1}\\epsilon\\,\\|u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}+\n\\big(c_{1}\\epsilon\\,|z|+c_{2}(1+\\epsilon^{-3})\\big)\\|R_{z}u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}\\Big)\\,.\n$$\nTo conclude the proof it suffices to show that \n$$\n\\|R_{z}\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\!,H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}}\\lesssim{d^{-\\gamma}_{z}}\\,,\n$$\nequivalently\n\\begin{equation}\\label{eqv}\n\\|R_{z}\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma),H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)}\\lesssim\n{d^{-\\gamma}_{z}}\\,,\n\\end{equation}\nwhere $\\gamma>0$ and $d_{z}$ is the distance of $z$ from $[0,+\\infty)$.\n\\par\nLet $u\\equiv u_{\\rm in}\\oplus u_{\\rm ex}\\in H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)$; then $u= 1_{\\Omega_{\\rm in}}\\widetilde u_{\\rm in}+1_{\\Omega_{\\rm ex}}\\widetilde u_{\\rm ex}$, where $\\widetilde u_{\\rm in\/\\rm ex}\\in H^{1}(\\mathbb{R}^{3})$ is such that $\\widetilde u_{\\rm in\/\\rm ex}|\\Omega_{\\rm in\/\\rm ex}=u_{\\rm in\/\\rm ex}$. One has\n\\begin{align*}\n&\\|R_{z}u\\|^{2}_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)}=\\|R_{z}u\\|^{2}_{H^{1}(\\mathbb{R}^{3})}\n=\\|R_{z}u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}+\n \\|\\nabla R_{z}u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\n =\\|R_{z}u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}+\n \\|R_{z}\\nabla u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\n\\\\\n\\le&\\|R_{z}u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}+\n\\|R_{z}(1_{\\Omega_{\\rm in}}\\nabla\\widetilde u_{\\rm in}+1_{\\Omega_{\\rm ex}}\\nabla\\widetilde u_{\\rm ex}\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\n+\\|R_{z}(\\widetilde u_{\\rm in}\\nabla 1_{\\Omega_{\\rm in}}+\\widetilde u_{\\rm ex}\\nabla 1_{\\Omega_{\\rm ex}}\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\\\\n\\le&\\|R_{z}\\|^{2}_{L^{2}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3})}\\|u\\|^{2}_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)}\n+\\|S\\!L_{z}\\nu[\\gamma_{0}]u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\,,\n\\end{align*}\nwhere $\\nu$ is the exterior normal to $\\Gamma$ and $S\\!L_{z}$ is the single-layer operator, i.e., $S\\!L_{z}:=(\\gamma_{0}R_{\\bar z})^{*}=R_{z}\\gamma_{0}^{*}$ and $[\\gamma_{0}]$ denotes the jump of the Dirichlet trace across $\\Gamma$. Here we made use the distributional derivative of $\\nabla 1_{\\Omega_{\\rm in\/\\rm ex}}$, which, by the divergence theorem, gives, for any test function $\\varphi$,\n$$\n\\nabla 1_{\\Omega_{\\rm in\/\\rm ex}}(\\varphi)=\\mp\\int_{\\Omega_{\\rm in\/\\rm ex}}\\nabla\\varphi(x)\n\\,dx=\\mp\\int_{\\Gamma}\\nu(x)\\gamma_{0}\\varphi(x)\\,d\\sigma(x)=\\mp\\gamma_{0}^{*}\\nu\\gamma_{0}\\varphi\\,.\n$$\nSince \n$$\n\\|S\\!L_{z}\\nu[\\gamma_{0}]\\|_{L^{2}(\\mathbb{R}^{3})}\\le\n\\|S\\!L_{z}\\|_{H^{1\/2}(\\Gamma),L^{2}(\\mathbb{R}^{3})}\n\\|[\\gamma_{0}]\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma),H^{1\/2}(\\Gamma)}\\|u\\|_{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)}\\,,\n$$\n\\begin{align*}\n&\\|S\\!L_{z}\\|_{H^{1\/2}(\\Gamma),L^{2}(\\mathbb{R}^{3})}\\le\\|S\\!L_{z}\\|_{H^{-1\/2}(\\Gamma),L^{2}(\\mathbb{R}^{3})}\n\\|S\\!L_{\\bar z}^{*}\\|_{L^{2}(\\mathbb{R}^{3}),H^{1\/2}(\\Gamma)}\\\\\n=&\n\\|\\gamma_{0}R_{z}\\|_{L^{2}(\\mathbb{R}^{3}),H^{1\/2}(\\Gamma)}\n\\le\\|\\gamma_{0}\\|_{H^{1}(\\mathbb{R}^{3}),H^{1\/2}(\\Gamma)}\n\\|R_{z}\\|_{L^{2}(\\mathbb{R}^{3}),H^{1}(\\mathbb{R}^{3})}\n\\end{align*}\nand\n$$\n\\|R_{z}\\|_{L^{2}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3})}\\le d_{z}^{-1}\\,,\n\\qquad\\|R_{z}\\|_{L^{2}(\\mathbb{R}^{3}),H^{1}(\\mathbb{R}^{3})}\\le d_{z}^{-1\/2}\\,,\n$$\n\\eqref{eqv} follows and the proof is concluded.\n\\end{proof}\n\\begin{remark} By Lemma \\ref{vH-1}, \n$$\n\\Lambda^{\\!\\mathsf v}_{z}|H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)=( 1- \\mathsf v R_{z}) ^{-1}\\mathsf v\\,.\n$$\n\\end{remark}\n\\begin{remark} By the same kind of proof (indeed a shorter one) as in Lemma \\ref{vH-1}, one gets \n$$\n( 1-\\mathsf v R_{z}) ^{-1}=\\sum_{k=0}^{+\\infty}\\left( \\mathsf v R_{z}\\right) ^{k}\n\\in{\\mathscr B}( H^{-1}(\\mathbb{R}^{3}))\n$$\nand, by \\eqref{v-sob}, $\\mathsf v\\in{\\mathscr B}( H^{1}(\\mathbb{R}^{3}), H^{-1}(\\mathbb{R}^{3}))$. Therefore, $( 1- \\mathsf v R_{z}) ^{-1}\\mathsf v\\in {\\mathscr B}( H^{1}(\\mathbb{R}^{3}), H^{-1}(\\mathbb{R}^{3}))$ and, by \\eqref{L-int} and \\eqref{RF-int},\n\\begin{equation}\\label{LvHs1}\n\\Lambda^{\\!\\mathsf v}_{z}|H^{s}(\\mathbb{R}^{3})=( 1- \\mathsf v R_{z}) ^{-1}\\mathsf v\\,,\\qquad 1\\le s\\le 2\\,.\n\\end{equation}\nBy duality, similarly to Remark \\ref{Lt}, $( 1-R_{z}\\mathsf v) ^{-1}\\in{\\mathscr B}( H^{1}(\\mathbb{R}^{3}))$ and \n\\eqref{LtL2} improves to\n\\begin{equation}\\label{LvHs2}\n\\Lambda^{\\!\\mathsf v}_{z}|H^{s}(\\mathbb{R}^{3})=\\mathsf v( 1- \\mathsf v R_{z}) ^{-1}\\,,\\qquad 0\\le s\\le 1\\,.\n\\end{equation}\n\\end{remark}\n\\subsection{\\label{Sec_Layer}Boundary layer operators.}\nWe introduce the interior\/exterior Dirichlet and Neumann trace operators\n$$\n\\gamma_{0}^{\\rm in\/\\rm ex}:H^{s+1\/2}(\\Omega_{\\rm in\/ex})\\to B_{2,2}^{s}(\\Gamma)\\,,\\qquad s>0\\,, \n$$\n$$\n\\gamma_{1}^{\\rm in\/\\rm ex}:H^{s+3\/2}(\\Omega_{\\rm in\/ex})\\to B_{2,2}^{s}(\\Gamma)\\,,\\qquad s>0\\,,\n$$\nwhere $\\Omega_{\\rm in}\\equiv\\Omega$ and $\\Omega_{\\rm ex}:=\\Omega_{\\rm ex}$. The Besov-like trace spaces $B_{2,2}^{s}( \\Gamma ) $ \nidentify with $H^{s}(\\Gamma) $ when $|s|\\le k+1$ and\n$\\Gamma$ is of class $\\mathcal{C}^{k,1}$ (see \\cite{JoWa}). Then, we define the bounded linear operators \n\\begin{equation}\\label{g0}\n\\gamma_{0}:H^{s+1\/2}(\\mathbb{R}^{3}\\backslash\\Gamma)\\to B_{2,2}^{s}(\\Gamma)\\,,\\quad\\gamma_{0}u:=\\frac12\\,(\\gamma_{0}^{\\rm in}(u|\\Omega_{\\rm in})+\\gamma_{0}^{\\rm ex}(u|\\Omega_{\\rm ex}))\\,,\\qquad s>0\\,,\n\\end{equation}\n\\begin{equation}\\label{g1}\n\\gamma_{1}:H^{s+3\/2}(\\mathbb{R}^{3}\\backslash\\Gamma)\\to B_{2,2}^{s}(\\Gamma)\\,,\\quad\\gamma_{1}u:=\\frac12\\,(\\gamma_{1}^{\\rm in}(u|\\Omega_{\\rm in})+\\gamma_{1}^{\\rm ex}(u|\\Omega_{\\rm ex}))\\,,\\qquad s>0\\,.\n\\end{equation}\nThe corresponding trace jump bounded operators are defined by\n\\begin{equation}\n[\\gamma_{0}]:H^{s+1\/2}( \\mathbb{R}^3\\backslash\\Gamma ) \\rightarrow B_{2,2}^{s}(\\Gamma)\\,,\\quad[\n\\gamma_{0}]u:=\\gamma_{0}^{\\-}(u|\\Omega_{\\rm in})-\\gamma_{0}^{\\+}(u|\\Omega_{\\rm ex})\\,,\n\\end{equation}%\n\\begin{equation}\n[\\gamma_{1}]:H^{s+3\/2}( \\mathbb{R}^3\\backslash\\Gamma )\\rightarrow B_{2,2}^{s}(\\Gamma)\\,,\\quad[\n\\gamma_{1}]u:=\\gamma_{1}^{\\-}(u|\\Omega_{\\rm in})-\\gamma_{1}^{\\+}(u|\\Omega_{\\rm ex})\\,.\n\\end{equation}\nBy \\cite[Lemma 4.3]{McL}, the trace maps $\\gamma_{1}^{\\-\/\\+}$ can be extended to the spaces $$H^{1}_{\\Delta}(\\Omega_{\\-\/\\+}):=\\{u_{\\-\/\\+}\\in H^{1}(\\Omega_{\\-\/\\+}):\\Delta_{\\Omega_{\\-\/\\+}}u_{\\-\/\\+}\\in L^{2}(\\Omega_{\\-\/\\+})\\}$$ \nas $H^{-1\/2}(\\Gamma)$-valued bounded operators:\n$$\n\\gamma_{1}^{\\-\/\\+}: H^{1}_{\\Delta}(\\Omega_{\\-\/\\+})\\to H^{-1\/2}(\\Gamma)\\,.\n$$\nThis gives the extensions of the maps $\\gamma_{1}$ and $[\\gamma_{1}]$ defined on $H^{1}_{\\Delta}(\\mathbb{R}^{3}\\backslash\\Gamma):=H^{1}_{\\Delta}(\\Omega_{\\-})\\oplus H^{1}_{\\Delta}(\\Omega_{\\+})$ with values in $H^{-1\/2}(\\Gamma)$. \\par\nThen, for any $z\\in\\mathbb{C}\\backslash(-\\infty,0]$,\none defines the single and double-layer operators\n\\begin{equation}\\label{SL}\nS\\!L_{z}:=(\\gamma_{0}R_{\\bar z})^{*}=R_{z}\\gamma_{0}^{*}\\in\\mathscr B(B_{2,2}^{-s}(\\Gamma),H^{3\/2-s}(\\mathbb{R}^{3}))\\,,\n\\qquad s>0\\,,\n\\end{equation}\n\\begin{equation}\\label{DL}\nD\\!L_{z}:=(\\gamma_{1}R_{\\bar z})^{*}=R_{z}\\gamma_{1}^{*}\\in\\mathscr B(B_{2,2}^{-s}(\\Gamma),H^{1\/2-s}(\\mathbb{R}^{3}))\\,,\\qquad s>0\\,.\n\\end{equation}\nBy \\eqref{g0}, one has\n\\begin{equation}\\label{BSL}\nS_{z}:=\\gamma_{0}S\\!L_{z}\\in \\mathscr B((H^{s-1\/2}(\\Gamma),H^{s+1\/2}(\\mathbb{R}^{3})))\\,,\\qquad -1\/2\\sup\\sigma(\\Delta+\\mathsf v)$ such that $[\\lambda_{\\mathsf v},+\\infty)\\subset Z^{\\circ}_{\\mathsf v,d}\\cap Z^{\\circ}_{\\mathsf v,n}$; moreover, $Z^{\\circ}_{\\mathsf v,d}\\cap Z^{\\circ}_{0,d}\\not=\\varnothing $, $Z^{\\circ}_{\\mathsf v,n}\\cap Z^{\\circ}_{0,n}\\not=\\varnothing $, and both $Z^{\\circ}_{\\mathsf v,d}$ and $Z^{\\circ}_{\\mathsf v,n}$ can be chosen to be symmetric with respect to the real axis.\n\\end{lemma}\n\\begin{proof} At first, let us notice that it suffices to show that the bounded inverses exist for any real $\\lambda \\ge \\lambda_{\\mathsf v}$ for some $\\lambda_{\\mathsf v}>\\sup\\sigma(\\Delta+\\mathsf v)$. Then, by the continuity of the maps $z\\mapsto S^{\\mathsf v}_{z}$ and $z\\mapsto D^{\\mathsf v}_{z}$, the bounded inverses exist in a complex open neighbourhood of $[\\lambda_{\\mathsf v},+\\infty)$.\\par\nWe proceed as in the proof of \\cite[Lemma 3.2]{JDE16}. By $(-(\\Delta+\\mathsf v)+\\lambda)S\\!L^{\\mathsf v}_{\\lambda}|\\Omega_{\\rm in\/\\rm ex}=0$, by Green's formula and by \\eqref{jumpv0}, one gets, for any $\\phi\\in H^{-1\/2}(\\Gamma)$,\n\\begin{align*}\n0=&\\|\\nabla S\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\n-\\langle \\mathsf v S\\!L^{\\mathsf v}_{\\lambda}\\phi,S\\!L^{\\mathsf v}_{\\lambda}\\phi\\rangle _{H^{-1}(\\mathbb{R}^{3}),H^{1}(\\mathbb{R}^{3})}\n+\\lambda\\,\\|S\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\\\\n&+\\langle [ \\gamma_{1}] S\\!L^{\\mathsf v}_{\\lambda}\\phi ,\\gamma_{0}S\\!L_{\\lambda}\\phi\\rangle _{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\\\\\n=&\\|\\nabla S\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\n-\\langle \\mathsf v S\\!L^{\\mathsf v}_{\\lambda}\\phi,S\\!L^{\\mathsf v}_{\\lambda}\\phi\\rangle _{H^{-1}(\\mathbb{R}^{3}),H^{1}(\\mathbb{R}^{3})}\n+\\lambda\\,\\|S\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\\\\n&-\\langle \\phi ,S^{\\mathsf v}_{\\lambda}\\phi\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\\,.\n\\end{align*}\nThen, by \\eqref{V_est},\n\\[\n\\langle \\phi ,\\gamma_{0}S^{\\mathsf v}_{\\lambda}\\phi\\rangle _{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\\geq( 1-c_{1}\\varepsilon) \\|\\nabla S\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\n+(\\lambda-c_{2}(1+\\varepsilon^{-3}))\\|S\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\,.\n\\]\nChoosing $\\varepsilon>0$ such that $c_{1}\\varepsilon<1$ and then $\\lambda\\in\\varrho(\\Delta+\\mathsf v)$ such that $\\lambda>c_{2}(1+\\varepsilon^{-3})$ (this is always possible since $\\Delta+\\mathsf v$ in bounded from above), one gets\n\\[\n\\langle \\phi,S_{\\lambda}^{\\mathsf v}\\phi\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\n\\gtrsim\\|S\\!L_{\\lambda}^{\\mathsf v}\\phi \\|_{H^{1}( \\mathbb{R}^{3}) }^{2}\\,.\n\\]\nBy Green's formula again, one has, for any $u_{\\rm in\/\\rm ex},v_{\\rm in\/\\rm ex}\\in H^{1}(\\Omega_{\\rm in\/\\rm ex})$, $(-(\\Delta+\\mathsf v)+\\lambda)u_{\\rm in\/\\rm ex}\\in L^{2}(\\Omega_{\\rm in\/\\rm ex})$,\n\\begin{align}\\label{Gf}\n&\\langle(-(\\Delta+\\mathsf v)+\\lambda)u_{\\rm in\/\\rm ex},v_{\\rm in\/\\rm ex}\\rangle_{L^{2},H^{1}(\\Omega_{\\rm in\/\\rm ex})}=\n\\langle\\nabla u_{\\rm in\/\\rm ex},\\nabla v_{\\rm in\/\\rm ex}\\rangle_{L^{2}(\\Omega_{\\rm in\/\\rm ex})}\\nonumber\\\\\n&-\\langle \\mathsf v u_{\\rm in\/\\rm ex},v_{\\rm in\/\\rm ex}\\rangle_{H^{-1}(\\Omega_{\\rm in\/\\rm ex}),\nH^{1}(\\Omega_{\\rm in\/\\rm ex})}+\\lambda\\,\\langle u_{\\rm in\/\\rm ex},v_{\\rm in\/\\rm ex}\\rangle_{L^{2}(\\mathbb{R}^{3})}\\\\\n&\n\\pm \\langle \\gamma_{1}^{\\rm in\/\\rm ex}u_{\\rm in\/\\rm ex},\\gamma^{\\rm in\/\\rm ex}_{0} v_{\\rm in\/\\rm ex}\n\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\\,.\\nonumber\n\\end{align}\nBy \n$$\n\\big|\\langle \\mathsf v u_{\\rm in\/\\rm ex},v_{\\rm in\/\\rm ex}\\rangle_{H^{-1}(\\Omega_{\\rm in\/\\rm ex}),\nH^{1}(\\Omega_{\\rm in\/\\rm ex})}\\big|\\lesssim \\|u_{\\rm in\/\\rm ex}\\|_{H^{1}(\\Omega_{\\rm in\/\\rm ex})}\n\\|v_{\\rm in\/\\rm ex}\\|_{H^{1}(\\Omega_{\\rm in\/\\rm ex})}\\,,\n$$\n\\eqref{Gf} gives,\n\\begin{align*}\n&\\big|\\langle \\gamma_{1}^{\\rm in\/\\rm ex}u_{\\rm in\/\\rm ex},\\gamma^{\\rm in\/\\rm ex}_{0} v_{\\rm in\/\\rm ex}\n\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\\big|\\\\\n\\lesssim\n&\\big(\\|u_{\\rm in\/\\rm ex}\\|_{H^{1}(\\Omega_{\\rm in\/\\rm ex})}+\\|(-(\\Delta+\\mathsf v)+\\lambda)u_{\\rm in\/\\rm ex}\\|_{H^{-1}(\\Omega_{\\rm in\/\\rm ex})}\\big)\n\\|v_{\\rm in\/\\rm ex}\\|_{H^{1}(\\Omega_{\\rm in\/\\rm ex})}\\,.\n\\end{align*}\nSince $\\gamma^{\\rm in\/\\rm ex}_{0}:H^{1}(\\Omega_{\\rm in\/\\rm ex})\\to H^{1\/2}(\\Gamma)$ is surjective, finally one gets \n\\begin{equation}\\label{cmp}\n\\|\\gamma_{1}^{\\rm in\/\\rm ex}u_{\\rm in\/\\rm ex}\\|_{H^{-1\/2}(\\Gamma)}\\lesssim \n\\|u_{\\rm in\/\\rm ex}\\|_{H^{1}(\\Omega_{\\rm in\/\\rm ex})}+\\|(-(\\Delta+\\mathsf v)+\\lambda)u_{\\rm in\/\\rm ex}\\|_{H^{-1}(\\Omega_{\\rm in\/\\rm ex})}\\,.\n\\end{equation}\nThen, proceeding as in \\cite[Lemma 3.2]{JDE16} (compare (3.31) there with \\eqref{cmp} here), this\nyields\n\\begin{equation*}\n\\langle \\phi,S_{\\lambda}^{\\mathsf v}\\phi\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\n\\gtrsim \\|\\phi\\|_{H^{-1\/2}(\\Gamma)}^{2}\\label{SL_coer}%\n\\end{equation*}\nand so $(S_{\\lambda}^{\\mathsf v}) ^{-1}\\in{\\mathscr B}( H^{1\/2}(\n\\Gamma ) ,H^{-1\/2}(\\Gamma) )$ by the Lax-Milgram theorem.\\par\nAs regards $D_{\\lambda}^{\\mathsf v}$, the proof is almost the same. By $(-(\\Delta+\\mathsf v)+\\lambda)D\\!L^{\\mathsf v}_{\\lambda}|\\Omega_{\\rm in\/\\rm ex}=0$, by Green's formula and by \\eqref{jumpv1}, one gets, for any $\\phi\\in H^{1\/2}(\\Gamma)$,\n\\begin{align*}\n0=&\\|\\nabla D\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\Omega_{\\rm in})}+\n\\|\\nabla D\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\Omega_{\\rm ex})}\n-\\langle \\mathsf v D\\!L^{\\mathsf v}_{\\lambda}\\phi,D\\!L^{\\mathsf v}_{\\lambda}\\phi\\rangle _{H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*},H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)}\n+\\lambda\\,\\|D\\!L^{\\mathsf v}_{\\lambda}\\phi\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\\\\n&+\\langle D^{\\mathsf v}_{\\lambda}\\phi ,\\phi\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\\,.\n\\end{align*}\nwhich leads to \n\\[\n-\\langle D_{\\lambda}^{\\mathsf v}\\phi,\\phi\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\n\\gtrsim\\|D\\!L_{\\lambda}^{\\mathsf v}\\phi \\|_{H^{1}( \\mathbb{R}^{3}\\backslash\\Gamma) }^{2}\\,.\n\\]\nThen, proceeding as in \\cite[Lemma 3.2]{JDE16}, by \\eqref{cmp}, this\nyields\n$$\n-\\langle D_{\\lambda}^{\\mathsf v}\\phi,\\phi\\rangle_{H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma)}\n\\gtrsim \\|\\phi\\|_{H^{1\/2}(\\Gamma)}^{2}\n$$\nand so $(D_{\\lambda}^{\\mathsf v}) ^{-1}\\in{\\mathscr B}( H^{-1\/2}(\n\\Gamma ) ,H^{1\/2}(\\Gamma) )$ by the Lax-Milgram theorem.\n\\end{proof}\n\n\n\\section{Laplacians with regular and singular perturbations}\nHere we apply the abstract results in Section \\ref{Sec_Krein}, presenting various examples were \nthe self-adjoint operator $A$ is the free Laplacian $\\Delta:H^{2}(\\mathbb{R}^{3})\\subset L^{2}(\\mathbb{R}^{3})\\to L^{2}(\\mathbb{R}^{3})$ and $A_{B_{1}}=\\Delta+\\mathsf v$. All over this section we consider a Kato-Rellich potential $\\mathsf v=\\mathsf v_{2}+\\mathsf v_{\\infty}$ of short-range type (however, see next Remark \\ref{suff}), i.e.,\n\\begin{equation}\\label{short}\n\\mathsf v_{2}\\in L^{2}(\\mathbb{R}^{3}),\\quad \\text{supp}(\\mathsf v_{2})\\ \\text{bounded}, \\qquad\n\\ |\\mathsf v_{\\infty}(x)|\\lesssim\\, (1+|x|\\,)^{-\\kappa(1+\\varepsilon)}\\,,\\quad\\kappa\\ge 1\\,,\\quad\\varepsilon>0\\,.\n\\end{equation}\nIn the next Lemmata we show that all our hypotheses but (H3) hold whenever $\\kappa=1$, while hypothesis (H3) holds whenever $\\kappa=2$; we conjecture that all our results hold true with $\\kappa=1$ and that the requirement $\\kappa=2$ is merely of technical nature. \\par\nWe take \n$$\n\\mathfrak h_{1}=H^{2}(\\mathbb{R}^{3})\\hookrightarrow \\mathfrak b_{1}=H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)\n\\hookrightarrow\\mathfrak h_{1}^{\\circ}=L^{2}(\\mathbb{R}^{3})\n\\,,\n$$\nand, introducing the multiplication operator $\\langle x\\rangle$ by $\\langle x\\rangle u: x\\mapsto(1+|x|^{2})^{1\/2}u(x)$, we define\n\\begin{equation}\\label{tau1}\n\\tau_{1}:H^{2}(\\mathbb{R}^{3})\\to H^{2}(\\mathbb{R}^{3})\\,,\\quad \\tau_{1}u:=\\langle x\\rangle^{-s}u\\,,\\qquad\n1<2s< 1+\\varepsilon\\,,\n\\end{equation}\nand\n$$\nB_{1}u:=\\langle x\\rangle^{2s}\\mathsf v u\\,.\n$$\nBy our hypotheses on $\\mathsf v$ and $s$, one has $\\langle x\\rangle^{2s}\\mathsf v\\in L^{2}(\\mathbb{R}^{3})+L^{\\infty}(\\mathbb{R}^{3})$ and so, by Lemma \\ref{v},\n$$\nB_{1}\\in\\mathscr B(H^{1}( \\mathbb{R}^{3}\\backslash\\Gamma),H^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,.\n$$\nConsidering the weight $\\varphi(x)=(1+|x|^{2})^{w\/2}$, $w\\in\\mathbb{R}$, we use the notation $L^{2}_{\\varphi}(\\mathbb{R}^{3})\\equiv L^{2}_{w}(\\mathbb{R}^{3})$; $H^{k}_{w}(\\mathbb{R}^{3})$, $H^{k}_{w}(\\mathbb{R}^{3}\\backslash\\Gamma)$ denotes the corresponding scales of weighted Sobolev spaces. \\par \nSince \n$$\n\\langle x\\rangle^{w}\\in\\mathscr B(H^{1}_{w'}(\\mathbb{R}^{3}\\backslash\\Gamma),H^{1}_{w'+w}(\\mathbb{R}^{3}\\backslash\\Gamma))$$ and, by duality, $$\\langle x\\rangle^{w}\\in\\mathscr B(H^{1}_{w'}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*},H^{1}_{w'-w}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,,$$ \none gets \n\\begin{equation}\\label{v-w}\n\\langle x\\rangle^{-w-2s}B_{1}\\langle x\\rangle^{w}=\\mathsf v\\in \\mathscr B(H^{1}_{w}(\\mathbb{R}^{3}\\backslash\\Gamma),H^{1}_{-w-2s}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,.\n\\end{equation} \nSince \n\\begin{equation}\\label{R-w}\nR_{z}\\in\\mathscr B(H^{-1}_{w}(\\mathbb{R}^{3}),H^{1}_{w}(\\mathbb{R}^{3}))\\hookrightarrow \\mathscr B(H^{1}_{-w}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*},H^{1}_{w}(\\mathbb{R}^{3}\\backslash\\Gamma))\\,,\n\\end{equation}\none has \n$$\\tau_{1}G^{1}_{z}=\\langle x\\rangle^{-s}R_{z}\\langle x\\rangle^{-s}\\in\\mathscr B(H_{-w}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*},H_{w+2s}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma))\\,.\n$$ \nIn particular, one has $\\tau_{1}G^{1}_{z}\\in\\mathscr B(H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}),H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma))$.\\par\nSince $(1-\\mathsf v R_{z})$ is invertible, \n$$\nM_{z}^{B_{1}}=1-B_{1}\\tau_{1}G^{1}_{z}=1-\\langle x\\rangle^{s}\\mathsf v R_{z}\\langle x\\rangle^{-s}=\n\\langle x\\rangle^{s}(1-\\mathsf v R_{z})\\langle x\\rangle^{-s\n$$ \nis invertible as well an\n\\begin{equation}\\label{LbLv}\n\\Lambda_{z}^{B_{1}}=(M_{z}^{B_{1}})^{-1}B_{1}=\\langle x\\rangle^{s}(1-\\mathsf v R_{z})^{-1}\\langle x\\rangle^{s}\\mathsf v=\\langle x\\rangle^{s}\\Lambda_{z}^{\\!\\mathsf v}\\langle x\\rangle^{s}\n\\,.\n\\end{equation}\n\\begin{lemma} Let $\\mathsf v$ be as in \\eqref{short}, with $\\kappa=1$. Then, for any $s$ such that $1< 2s<1+\\varepsilon$,\n$$\n\\Lambda_{z}^{\\!\\mathsf v}\n\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma),H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,.\n$$\n\\end{lemma}\n\\begin{proof} By \\eqref{v-w} and by $(1-R_{z}\\mathsf v)^{-1}\\in \\mathscr B(H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*})$ (see Lemma \\ref{vH-1}), one has\n$$\nR_{z}\\Lambda^{\\mathsf v}_{z}=(1-\\mathsf v R_{z}\\mathsf v)^{-1}\\mathsf v\\in \\mathscr B(H_{-2s}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma), H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma))\\,.\n$$\nBy interpolation (using \\cite{CE}), \n$$\nR_{z}\\Lambda^{\\mathsf v}_{z}\\in \\mathscr B(H_{-s}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma), H_{s}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma))\n$$\nand so \n$$\n\\Lambda^{\\mathsf v}_{z}\\in \\mathscr B(H_{-s}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma), H_{-s}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,.\n$$\n\\end{proof}\nBy the previous Lemma and \\eqref{LbLv}, \n$$\n\\Lambda_{z}^{B_{1}}\\in{\\mathscr B}(H^{1}( \\mathbb{R}^{3}\\backslash\\Gamma),H^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*})\\,.\n$$ \nThen, by Theorem \\ref{KR} and Corollary \\ref{cor1} (see Remark \\ref{tpt}), $Z_{B_{1}}=\\varrho(\\Delta+\\mathsf v)\\cap\\mathbb{C}\\backslash(-\\infty,0]$ and\n\\begin{align*}\n(-A_{B_{1}}+z)^{-1}=&R_{z}^{B_{1}}=R_{z}+R_{z}\\langle x\\rangle^{-s}\\Lambda_{z}^{B_{1}}\\langle x\\rangle^{-s}R_{z}\\qquad z\\in\\varrho(\\Delta+\\mathsf v)\\cap\\mathbb{C}\\backslash(-\\infty,0]\\\\\n=&\nR_{z}+R_{z}\\Lambda_{z}^{\\mathsf v}R_{z}\\\\\n=&R^{\\mathsf v}_{z}=(-(\\Delta+\\mathsf v)+z)^{-1}\\,.\n\\end{align*}\n\\begin{remark}\\label{suff} By the above relation, the self-adjoint operator provided in Corollary \\ref{cor1} by the choice $\\tau_{1}u=\\langle x\\rangle^{-s}u$, $B_{1}=\\langle x\\rangle^{2s}\\mathsf v$, coincides with the one corresponding to $\\tau_{1}u=u$, $B_{1}=\\mathsf v$. The first choice is dictated by the need to obtain LAP and a representation formula for the scattering couple $(A_{\\mathsf B}, A)$; whenever one is only interested in providing a resolvent formula for $A_{\\mathsf B}$, then the second choice is preferable: in this case, one does not need to work with a short-range potential $\\mathsf v$: a Kato-Rellich potential suffices.\n\\end{remark}\nNow, we prove the validity of the hypotheses (H1)-(H7); here and below we refer to the weighted spaces $L_{s}^{2}(\\mathbb{R}^{3})$ with $1<2s<1+\\varepsilon$. \\par\n\\begin{lemma} Let $\\mathsf v$ be short-range as in \\eqref{short}, with $\\kappa=1$.\nThen hypotheses (H1), (H2), (H6), (H7.1), (H7.2), (H7.3) hold true. \n\\end{lemma}\n\\begin{proof} By \\cite[Lemma 1, page 170]{ReSi IV}, $R_{z}=(-\\Delta+z)^{-1}\\in\\mathscr B(L^{2}_{s}(\\mathbb{R}^{3}))$ for any $z\\in \\mathbb{C}\\backslash(-\\infty,0]$. Therefore, by the resolvent identity $R^{\\mathsf v}_{z}=R_{z}(1-\\mathsf v R^{\\mathsf v}_{z})$, $z\\in\\varrho(-(\\Delta+\\mathsf v))$, and by $R^{\\mathsf v}_{z}\\in \\mathscr B(L_{s}^{2}(\\mathbb{R}^{3}), H^{2}(\\mathbb{R}^{3}))$, hypothesis (H1) is consequence of $\\mathsf v=\\mathsf v_{2}+\\mathsf v_{\\infty}\\in\\mathscr B(H^{2}(\\mathbb{R}^{3}),L_{s}^{2}(\\mathbb{R}^{3}))$. Since $\\mathsf v_{2}$ has a compact support, $\\mathsf v_{2}\\in \\mathscr B(H^{2}(\\mathbb{R}^{3}), L_{s}^{2}(\\mathbb{R}^{3}))$ by Lemma \\ref{v}. As regards $\\mathsf v_{\\infty}$, one has\n\\begin{align*}\n\\|\\mathsf v_{\\infty} u\\|^{2}_{L_{s}^{2}(\\mathbb{R}^{3})}=\\int_{\\mathbb{R}^{3}}|\\mathsf v_{\\infty} u|^{2}(1+|x|^{2})^{s}dx\\le c\\int_{\\mathbb{R}^{3}}(1+|x|)^{-2(1+\\varepsilon)}(1+|x^{2}|)^{s}\n|u|^{2}dx\n\\le \\,c\\,\\|u\\|^{2}_{L^{2}(\\mathbb{R}^{3})}\\,.\n\\end{align*} \nBy \\cite[Theorem 4.1]{Agmon}, LAP holds for $A=\\Delta$; hence (H7.1) is satisfied. By the sort-range hypothesis on $\\mathsf v$ and by \\cite[Theorem 4.2]{Agmon}, LAP holds for $A_{B_{1}}\\equiv\\Delta+\\mathsf v$ as well and, by \\cite[Theorems 6.1 and 7.1]{Agmon} asymptotic completeness holds for the scattering couple $(A_{B_{1}},A)\\equiv(\\Delta+\\mathsf v,\\Delta)$. Hence hypotheses (H1), (H2)\\footnote{ here ${e}(A_{B_{1}})$ is a discrete set in $(-\\infty,0)$ by \\cite[Theorem 4.2]{Agmon}; it is given by the (possibly empty) set of negative (embedded) eigenvalues of $\\Delta+\\mathsf v$. It is known that there are no negative eigenvalue either if $|\\mathsf v(x)|\\lesssim (1+|x|)^{-(1+\\varepsilon)}$ (see \\cite{Vak}) or if $\\mathsf v\\in L^{3\/2}(\\mathbb{R}^{3})$, i.e., if $\\kappa=2$ in \\eqref{short} (see \\cite{IJ}).} and (H6) are verified. \\par \nBy $R_{z}\\in\\mathscr B(L^{2}_{-s}(\\mathbb{R}^{3}),H^{2}_{-s}(\\mathbb{R}^{3}))$, one gets $G_{z}^{1*}=\\langle x\\rangle^{-s}R_{z}\\in\\mathscr B(L^{2}_{-s}(\\mathbb{R}^{3}),H^{2}(\\mathbb{R}^{3}))$ and so, by duality, $G_{z}^{1}\\in\\mathscr B(H^{-2}(\\mathbb{R}^{3}), L^{2}_{s}(\\mathbb{R}^{3}))$; moreover, by $R^{\\pm}_{\\lambda}\\in\\mathscr B(L^{2}_{s}(\\mathbb{R}^{3}),H^{2}_{-s}(\\mathbb{R}^{3}))$ and by a similar duality argument, one gets $G_{\\lambda}^{1,\\pm}\\in\\mathscr B(H^{-2}(\\mathbb{R}^{3}), L^{2}_{-s}(\\mathbb{R}^{3}))$. Thus hypothesis (H7.2) holds.\\par \nBy $\\Lambda_{z}^{\\!B_{1}}=\\langle x\\rangle^{s}\\Lambda_{z}^{\\!\\mathsf v}\\langle x\\rangle^{s}$, hypothesis (H7.3) is equivalent to the existence in $\\mathscr B(H_{-s}^{2}(\\mathbb{R}^{3}),H_{s}^{-2}(\\mathbb{R}^{3}))$ of $\\lim_{\\epsilon\\searrow 0}\\Lambda_{\\lambda\\pm i\\epsilon}^{\\!\\mathsf v}=\\lim_{\\epsilon\\searrow 0}(1-\\mathsf v R_{\\lambda\\pm i\\epsilon})^{-1}\\mathsf v$. By \\eqref{short}, $\\mathsf v\\in \\mathscr B(H_{-s}^{2}(\\mathbb{R}^{3}),L_{s}^{2}(\\mathbb{R}^{3}))$. Then, $\\lim_{\\epsilon\\searrow 0}(1-\\mathsf v R_{\\lambda\\pm i\\epsilon})^{-1}$ exists in $\\mathscr B(L^{2}_{s}(\\mathbb{R}^{3}))$ (see \\cite[proof of Theorem XIII.33, page 177]{ReSi IV}) and so (H7.3) holds.\n\\end{proof}\n\\begin{lemma}\\label{5.4} Let $\\mathsf v$ be short-range as in \\eqref{short}, with $\\kappa=2$.\nThen hypothesis (H3) holds true. \n\\end{lemma}\n\\begin{proof} The proof is the same as the one for \\cite[Lemma 4.5]{JDE18}, once one proves that \n\\begin{equation}\\label{once}\n\\mathsf v R^{\\mathsf v,\\pm}_{\\lambda}\\in\\mathscr B(L^{2}_{2s}(\\mathbb{R}^{3}))\\,.\n\\end{equation}\nSince $R^{\\mathsf v,\\pm}_{\\lambda}\\in\\mathscr B(L^{2}_{2s}(\\mathbb{R}^{3}), H^{2}_{-2s}(\\mathbb{R}^{3}))$, \n\\eqref{once} is consequence of \n\\begin{equation}\\label{vw}\n\\mathsf v=\\mathsf v_{2}+\\mathsf v_{\\infty}\\in \\mathscr B(H^{2}_{-2s}(\\mathbb{R}^{3}), L^{2}_{2s}(\\mathbb{R}^{3}))\\,.\n\\end{equation}\nLemma \\ref{v} entails $\\mathsf v_{2}\\in \\mathscr B(H^{2}(\\mathbb{R}^{3}), L^{2}(\\mathbb{R}^{3}))$ and so, since $\\mathsf v_{2}$ has a compact support, one gets that $\\mathsf v_{2}$ satisfies \\eqref{vw}. As regards $\\mathsf v_{\\infty}$, one has, by $1<2s<1+\\varepsilon$,\n\\begin{align*}\n\\|\\mathsf v_{\\infty} u\\|^{2}_{L^{2}_{2s}(\\mathbb{R}^{3})}=&\\int_{\\mathbb{R}^{3}}|\\mathsf v_{\\infty} u|^{2}(1+|x|^{2})^{2s}dx\\le c\\int_{\\mathbb{R}^{3}}(1+|x|)^{-4(1+\\varepsilon)}(1+|x|^{2})^{4s}\n|u|^{2}(1+|x|^{2})^{-2s}dx\\\\\n\\le& \\,c\\,\\|u\\|^{2}_{L^{2}_{-2s}(\\mathbb{R}^{3})}\\le c\\,\\|u\\|^{2}_{H^{2}_{-2s}(\\mathbb{R}^{3})}\n\\end{align*} \nand so $\\mathsf v_{\\infty}$ satisfies \\eqref{vw} as well.\n\\end{proof}\nIn order to check the validity of the remaining hypotheses, we need to specify the map \n$\\tau_{2}$. In the following examples, according to the case, we take either \n\\begin{equation}\\label{td}\n\\tau_{2}=\\gamma_{0}:H^{2}(\\mathbb{R}^{3})\\to\\mathfrak h_{2}=B^{3\/2}_{2,2}(\\Gamma)\\hookrightarrow\\mathfrak b_{2}=H^{s_{\\circ}}(\\Gamma)\\,,\\quad 00$, one gets $\\Lambda^{\\!\\mathsf v}_{z}\\in \\mathscr B(L^{2}_{-(2s+\\gamma)}(\\mathbb{R}^{3}),L^{2}(\\mathbb{R}^{3}))$. Hence, by the resolvent formula \\eqref{RF-int} and by $R_{z}\\in \\mathscr B(L^{2}_{-(2s+\\gamma)}(\\mathbb{R}^{3}),H^{2}_{-(2s+\\gamma)}(\\mathbb{R}^{3}))$, one gets $R_{z}^{\\mathsf v}\\in \\mathscr B(L^{2}_{-(2s+\\gamma)}(\\mathbb{R}^{3}),H^{2}_{-(2s+\\gamma)}(\\mathbb{R}^{3}))$. This entails $\\gamma_{0}R_{z}^{B_{1}}=\\gamma_{0}R_{z}^{\\mathsf v}=\\gamma_{0}\\chi R_{z}^{\\mathsf v}\\in \\mathscr B(L^{2}_{-(2s+\\gamma)}(\\mathbb{R}^{3}),B^{2}_{2,2}(\\Gamma))$ and $\\gamma_{1}R_{z}^{B_{1}}=\\gamma_{1}R_{z}^{\\mathsf v}=\\gamma_{1}\\chi R_{z}^{\\mathsf v}\\in \\mathscr B(L^{2}_{-(2s+\\gamma)}(\\mathbb{R}^{3}),H^{1\/2}(\\Gamma))$. Then, by duality, one gets $G_{z}^{B_{1}}\\in \\mathscr B(\\mathfrak h_{2}^{*},L^{2}_{2s+\\gamma}(\\mathbb{R}^{3}))$. This shows that (H4.2) holds. \\par\nBy \\cite[Theorem 4.2]{Agmon}, the map $(\\mathbb{R}\\backslash{e}(A_{B_{1}}))\\cup\\mathbb{C}_{\\pm}\\ni z\\mapsto R^{B_{1},\\pm}_{z}=R_{z}^{\\mathsf v,\\pm}\\in \\mathscr B(L^{2}_{s}(\\mathbb{R}^{3}),H^{2}_{-s}(\\mathbb{R}^{3}))$ is continuos. Hence, $z\\mapsto\\gamma_{0}R^{B_{1},\\pm}_{z}=\\gamma_{0}R_{z}^{\\mathsf v,\\pm}=\\gamma_{0}\\chi R_{z}^{\\mathsf v,\\pm}$ and $z\\mapsto\\gamma_{1}R^{B_{1},\\pm}_{z}=\\gamma_{1}R_{z}^{\\mathsf v,\\pm}=\\gamma_{1}\\chi R_{z}^{\\mathsf v,\\pm}$ are continuos as $\\mathscr B(L^{2}_{s}(\\mathbb{R}^{3}),B^{3\/2}_{2,2}(\\Gamma))$-valued and $\\mathscr B(L^{2}_{s}(\\mathbb{R}^{3}),H^{1\/2}(\\Gamma))$-valued maps respectively. Then, by duality, $z\\mapsto G^{B_{1},\\pm}$ is continuos on $(\\mathbb{R}\\backslash{e}(A_{B_{1}}))\\cup\\mathbb{C}_{\\pm}$ as a $\\mathscr B(\\mathfrak h_{2}^{*}, L^{2}_{-s}(\\mathbb{R}^{3}))$-valued map. Since both $\\gamma_{0}:H^{2}(\\mathbb{R}^{2})\\to B^{3\/2}_{2,2}(\\Gamma)$ and $\\gamma_{1}:H^{2}(\\mathbb{R}^{2})\\to H^{1\/2}(\\Gamma)$ are surjective, $G^{B_{1},\\pm}_{z}\\in \\mathscr B(\\mathfrak h_{2}^{*}, L^{2}_{-s}(\\mathbb{R}^{3}))$ is the adjoint of a surjective map and hence is injective. Thus we proved that (H5) holds.\n\\end{proof}\nIn conclusion, we proved that all our hypotheses, except (H4.1), hold true without the need to specify the operators $B_{0}$ and $B_{2}$. The validity of hypothesis \n(H4.1), i.e. the semi-boundedness of $A_{\\mathsf B}$, will be checked case by case in the next examples. Before turning to such examples, we give a result that makes explicit the map $\\mathcal L_{\\lambda}$ appearing in Theorem \\ref{S-matrix}. \n\\begin{lemma}\\label{Llambda} Let $\\tau_{1}$ be as in \\eqref{tau1} and let $\\tau_{2}$ be either as in \\eqref{td} or as in \\eqref{tn}. \nThen $$L_{\\lambda}:=-\\mathcal L_{\\lambda}\\langle x\\rangle^{s}$$\nis $L^{2}({\\mathbb S}^{2})$-valued, where ${\\mathbb S}^{2}$ denotes the 2-dimensional unitary sphere in $\\mathbb{R}^{3}$, and\n$$\nL_{\\lambda}(u\\oplus\\phi)=\\frac{|\\lambda|^{\\frac{1}4}}{2^{\\frac12}}\\,\\left(L^{1}_{\\lambda}u+L^{2}_{\\lambda}\\phi\\right)\\,,\n$$\n$$\nL^{1}_{\\lambda}u(\\xi):=\\widehat{u}\\,(|\\lambda|^{1\/2}\\xi)\n$$\n$$\nL^{2}_{\\lambda}\\phi(\\xi):=\\frac1{(2\\pi)^{\\frac32}}\\,\\begin{cases}\\langle u^{\\xi}_{\\lambda}|\\Gamma,\\phi\\rangle_{H^{1\/2}(\\Gamma),H^{-1\/2}(\\Gamma)}\\,,&\\tau_{2}=\\gamma_{0}\\,,\\\\\ni\\,|\\lambda|^{\\frac12}\\nu\\!\\cdot\\! \\xi\\,\\langle u^{\\xi}_{\\lambda}|\\Gamma,\\phi\\rangle_{L^{2}(\\Gamma)}\\,,&\\tau_{2}=\\gamma_{1}\\,.\n\\end{cases}\n$$\nHere $u^{\\xi}_{\\lambda}$ is the plane wave with direction $\\xi\\in{\\mathbb S}^{2}$ and wavenumber $|\\lambda|^{\\frac12}$, i.e., $u^{\\xi}_{\\lambda}(x)=e^{i\\,|\\lambda|^{\\frac12}\\xi\\cdot x}$, $\\nu$ is the normal to $\\Gamma$ and $\\widehat u$ denotes the Fourier transform.\n\\end{lemma} \n\\begin{proof} The unitary \nmap $F:L^{2}(\\mathbb{R}^{n})\\to \\int^{\\oplus}_{(-\\infty,0)}L^{2}({\\mathbb S}^{2})\\,d\\lambda\\equiv L^{2}((-\\infty,0);L^{2}({\\mathbb S}^{2}))$ diagonalizing $A=\\Delta$ is given by \n\\begin{equation}\\label{fourier}\n(Fu)_\\lambda(\\xi):=-\\frac{|\\lambda|^{\\frac{1}4}}{2^{\\frac12}}\\, \\widehat u(|\\lambda|^{1\/2}\\xi)\\,.\n\\end{equation} \nTherefore, by $(\\mu-\\lambda)\\widehat{ R_{\\mu}f}(|\\lambda|^{1\/2}\\xi)=-\\widehat f(|\\lambda|^{1\/2}\\xi)$,\none gets \n\\begin{align*}\n(\\mu-\\lambda)(FR_{\\mu}\\tau_{1}^{*}\\langle x\\rangle^{s}u)_{\\lambda}(\\xi)=-\n\\frac{|\\lambda|^{\\frac{1}4}}{2^{\\frac12}}\\,\\widehat{u}\\,(|\\lambda|^{1\/2}\\xi)\n\\,.\n\\end{align*}\nAs regards $L^{2}_{\\lambda}$, the computation was given in \\cite[Theorem 5.1]{JMPA}. \n\\end{proof}\n\\begin{remark} Let us notice that, whenever $u\\in L^{2}_{s'}(\\mathbb{R}^{3})$, $s'>3\/2$,\n$$\nL^{1}_{\\lambda}u(\\xi)=\\frac1{(2\\pi)^{\\frac32}}\\,\\langle u^{\\xi}_{\\lambda},u\\rangle_{L^{2}_{-s'}(\\mathbb{R}^{3}),L^{2}_{s'}(\\mathbb{R}^{3})}\n$$\nand so $L^{1}_{\\lambda}$ and $L^{2}_{\\lambda}$ have a similar structure.\n\\end{remark}\n\\subsection{\\label{Sec_delta} Short-range potentials and semi-transparent boundary conditions\nof $\\delta_{\\Gamma }$-type} \nHere we take \n$$\n\\mathfrak h_{2}= B^{3\/2}_{2,2}(\\Gamma)\\hookrightarrow\\mathfrak b_{2}=\\mathfrak b_{2,2}= H^{s_{\\circ}}(\\Gamma)\n\\hookrightarrow\\mathfrak h_{2}^{\\circ}=L^{2}(\\Gamma)\\,,\\quad 0< s_{\\circ}<1\/2\\,,\n$$\n$$\n\\tau_{2}=\\gamma_{0}:H^{2}(\\mathbb{R}^{3})\\to B^{3\/2}_{2,2}(\\Gamma)\\,,\\qquad\nB_{0}=1\\,\n\\qquad B_{2}=\\alpha\\,,\n$$\nwhere \n$$\n\\alpha\\in\\mathscr B(H^{s_{\\circ}}(\\Gamma),H^{-s_{\\circ}}(\\Gamma))\\,,\\quad \\alpha^{*}=\\alpha\\,.\n$$ \nLet us notice (see \\cite[Remark 2.6]{JDE18}) that in the case $\\alpha$ is the multiplication operator associated to a real-valued function \n$\\alpha$, then $\\alpha\\in L^{p}(\\Gamma)$, $p>2$, fulfills our hypothesis. \\par\nFor any $z\\in\\mathbb{C}\\backslash(-\\infty,0]$, one has\n\\begin{equation}\nM_{z}^{\\mathsf B\n=1-\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & \\alpha%\n\\end{bmatrix}\n\\begin{bmatrix}\n\\langle x\\rangle^{-s}R_{z}\\langle x\\rangle^{-s} & \\langle x\\rangle^{-s}R_{z}\\gamma_{0}^{*}\\\\\n\\gamma_{0}R_{z} \\langle x\\rangle^{-s}& \\gamma_{0}R_{z}\\gamma_{0}^{*}%\n\\end{bmatrix}\n=\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{M}_{z}^{\\mathsf v,\\alpha}\\begin{bmatrix}\\langle x\\rangle^{-s}&0\\\\0&1\\end{bmatrix}\\,,\n\\end{equation}\n\\begin{equation}\n{M}_{z}^{\\mathsf v,\\alpha}:=%\n\\begin{bmatrix}\n1-\\mathsf v R_{z} & -\\mathsf v S\\!L_{z}\\\\\n-\\alphaS\\!L_{\\bar z}^{*} & 1-\\alpha S_{z}%\n\\end{bmatrix}\n\\,\n\\end{equation}\nBy the mapping properties provided in Sections \\ref{Sec_V} and \\ref{Sec_Layer}, by \\eqref{v-w} and \\eqref{R-w} with $w=-s$, one gets\n$$\nM_{z}^{\\mathsf v,\\alpha}\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{-s_{\\circ}}(\\Gamma))\\,.\n$$\nAccording to \n\\cite[Lemma 5.8]{JMPA}, for any $z\\in \\mathbb{C}\\backslash((-\\infty,0]\\cup\\sigma_{\\alpha})$, where $\\sigma_{\\alpha}\\subset (0,+\\infty)$ is finite, one has\n\\begin{equation}\n(M_{z}^{B_{0},B_{2}})^{-1}=(M_{z}^{\\alpha})^{-1}:=\n( 1-\\alpha S_{z}) ^{-1}\\in{\\mathscr B}( H^{-s_{\\circ}}(\n\\Gamma ) ) \\,.\\label{S-alpha}%\n\\end{equation}\nThus $$\nZ_{B_{0},B_{2}}=Z_{\\alpha}:=\\{z\\in \\mathbb{C}\\backslash(-\\infty,0]:(M_{w}^{\\alpha})^{-1}\\in{\\mathscr B}( H^{-s_{\\circ}}(\n\\Gamma ) ) ,\\ w=z,\\bar z\\}\\supseteq \\mathbb{C}\\backslash((-\\infty,0]\\cup\\sigma_{\\alpha})\n$$ \nand \n$$ \n\\Lambda_{z}^{B_{0},B_{2}}=(M_{z}^{B_{0},B_{2}})^{-1}B_{2}=\\Lambda_{z}^{\\!\\alpha}:=(1-\\alpha S_{z})^{-1}\\alpha\\in{\\mathscr B}( H^{s_{\\circ}}(\n\\Gamma ) , H^{-s_{\\circ}}(\\Gamma ) ) \\,.\n$$\nBy \\cite[Corollary 2.4]{JDE18}, for any $z\\in \\varrho(\\Delta+\\mathsf v)\\backslash\\sigma_{\\mathsf v,\\alpha}$, where $\\sigma_{\\mathsf v,\\alpha}\\subset\\mathbb{R}$ is finite, \n\\begin{equation}\n(\\widehat M_{z}^{\\mathsf B})^{-1}=(\\widehat M_{z}^{\\mathsf v,\\alpha})^{-1}:=\n( 1-\\alpha S_{z}^{\\mathsf v}) ^{-1}\n\\in{\\mathscr B}( H^{-s_{\\circ}}(\\Gamma ) ) \n\\,.\\label{Sv-alpha}%\n\\end{equation}\nThus \n$$\n\\widehat Z_{\\mathsf B}=\\widehat Z_{\\mathsf v,\\alpha}:=\\{z\\in \\varrho(\\Delta+\\mathsf v):(\\widehat M_{w}^{\\mathsf v,\\alpha})^{-1}\\in{\\mathscr B}( H^{-s_{\\circ}}(\n\\Gamma ) ) ,\\ w=z,\\bar z\\}\\supseteq \\varrho(\\Delta+\\mathsf v)\\backslash\\sigma_{\\mathsf v,\\alpha}\n$$ \nand\n$$\n\\widehat\\Lambda^{\\mathsf B}_{z}=(\\widehat M_{z}^{\\mathsf B})^{-1}B_{2}=\n\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}:=(1-\\alpha S^{\\mathsf v}_{z} )^{-1}\\alpha\n\\in{\\mathscr B}(H^{s_{\\circ}}(\\Gamma ), H^{-s_{\\circ}}(\\Gamma ) )\\,.\n$$\nHence, \n$$\n\\Lambda_{z}^{\\mathsf B}=\n\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}({M}_{z}^{\\mathsf v,\\alpha})^{-1}\\begin{bmatrix}\\langle x\\rangle^{-s}&0\\\\0&1\\end{bmatrix}\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & \\alpha%\n\\end{bmatrix}\n=\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{\\Lambda}_{z}^{\\!\\mathsf v,\\alpha}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\\,,\n$$\nwhere, by \\eqref{LB-new} and \\eqref{LB-new2},\n\\begin{align*}\n{\\Lambda}_{z}^{\\!\\mathsf v,\\alpha}:=&\n\\begin{bmatrix}\n\\Lambda_{z}^{\\!\\mathsf v}+\\Lambda_{z}^{\\!\\mathsf v}S\\!L_{z}\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}S\\!L_{\\bar z}^{*}\\Lambda_{z}^{\\!\\mathsf v}& \\Lambda_{z}^{\\!\\mathsf v}S\\!L_{z}\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}\n\\\\\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}S\\!L_{\\bar z}^{*}\\Lambda_{z}^{\\!\\mathsf v} & \n\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}\n\\end{bmatrix}\\\\\n=&\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\nS\\!L_{z}\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}S\\!L_{\\bar z}^{*}& S\\!L_{z}\n\\\\ S\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}\\end{bmatrix}\n\\,.\n\\end{align*}\nOne has \n\\begin{equation}\\label{Lbbvalpha}\n{\\Lambda}_{z}^{\\!\\mathsf v,\\alpha}\n\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)\\oplus H^{s_{\\circ}}(\\Gamma),H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{-s_{\\circ}}(\\Gamma))\\,.\n\\end{equation}\nBy Theorems \\ref{Th_Krein} and \\ref{Th-alt-res}, there follows \n\\begin{align}\nR_{z}^{\\mathsf v,\\alpha}=&R_{z}+\n\\begin{bmatrix}R_{z}\\langle x\\rangle^{-s}&S\\!L_{z}\\end{bmatrix}\n\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{\\Lambda}_{z}^{\\!\\mathsf v,\\alpha}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\n\\begin{bmatrix}\\langle x\\rangle^{2s}\\mathsf v \\langle x\\rangle^{-s}R_{z}\\\\ \\alphaS\\!L_{\\bar z}^{*}\\end{bmatrix}\n\\label{Rv-alpha-0}\n\\\\ \n=&R_{z}+\n\\begin{bmatrix}R_{z}&S\\!L_{z}\\end{bmatrix}\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\nS\\!L_{z}\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}S\\!L_{\\bar z}^{*}& S\\!L_{z}\n\\\\ S\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\alpha}\\end{bmatrix}\n\\begin{bmatrix}R_{z}\\\\ S\\!L_{\\bar z}^{*}\\end{bmatrix}\n\\label{Rv-alpha-1}\\\\ \n=&R_{z}^{\\mathsf v}+S\\!L_{z}^{\\mathsf v}{\\widehat\\Lambda}_{z}^{\\mathsf v,\\alpha}{S\\!L_{\\bar z}^{\\mathsf v}}^{*}\\,.\n\\label{Rv-alpha-2}\n\\end{align}\nis the resolvent of a self-adjoint operator $\\Delta^{\\!\\mathsf v,\\delta,\\alpha}$; \\eqref{Rv-alpha-0} holds for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,\\delta,\\alpha})\\cap\\mathbb{C}\\backslash(-\\infty,0]$, both\n\\eqref{Rv-alpha-1} and \\eqref{Rv-alpha-2} hold for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,\\delta,\\alpha})\\cap\\varrho(\\Delta+\\mathsf v)$. \n\\par\nBy Theorem \\ref{Th-add},\n$$\n\\Delta^{\\!\\mathsf v,\\delta,\\alpha}u=\\Delta u+\\mathsf v u+(\\alpha\\gamma_{0}u)\\delta_{\\Gamma}\\,.\n$$\nBy \\eqref{Rv-alpha-2} and by the mapping properties of $S\\!L^{\\mathsf v}_{z}$, one has \n$$\n\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,\\delta,\\alpha})\\subseteq H^{3\/2-s_{\\circ}}(\\mathbb{R}^{3})\\,.\n$$ \nMoreover, by $R^{\\mathsf v}_{z}u\\in H^{2}(\\mathbb{R}^{3})$, so that $[\\gamma_{1}]R^{\\mathsf v}_{z}u=0$, and by \\eqref{jumpv0}, one gets $[\\gamma_{1}]R^{\\mathsf v,\\alpha}_{z}u=-{\\widehat\\Lambda}_{z}^{\\mathsf v,\\alpha}{S\\!L_{\\bar z}^{\\mathsf v}}^{*}u=-\\widehat\\rho_{\\mathsf B}(R^{\\mathsf v,\\alpha}_{z}u)$. Hence, by Lemma \\ref{alt-abc}, \n$$\nu\\in\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,\\delta,\\alpha})\\quad\\Longrightarrow\\quad\\alpha\\gamma_{0}u+[\\gamma_{1}]u=0\\,.\n$$\nSince $\\widehat Z_{\\mathsf v,\\alpha}$ contains a positive half-line, $\\Delta^{\\!\\mathsf v,\\delta,\\alpha}$ is bounded from above and hypothesis (H4.1) holds. The scattering couple $(\\Delta^{\\!\\mathsf v,\\delta,\\alpha},\\Delta)$ is asymptotically complete and the corresponding scattering matrix is given by \n$$\n{\\mathcal S}_{\\lambda}^{\\mathsf v,\\alpha}=1-2\\pi iL_{\\lambda}\\Lambda^{\\mathsf v,\\alpha,+}_{\\lambda}L_{\\lambda}^{*}\\,,\\quad \\lambda\\in(-\\infty,0]\\backslash(\\sigma^{-}_{p}(\\Delta+\\mathsf v)\\cup \\sigma^{-}_{p}(\\Delta^{\\!\\mathsf v,\\delta,\\alpha}))\\,,\n$$\nwhere $L_{\\lambda}$ is given in Corollary \\ref{Llambda} and $\\Lambda^{\\mathsf v,\\alpha,+}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\Lambda^{\\mathsf v,\\alpha}_{\\lambda+i\\epsilon}$. This latter limit exists by Lemma \\ref{rmH7}; in particular, by \\eqref{LBpm2}, \n\\begin{align*}\n\\Lambda^{\\mathsf v,\\alpha,+}=&\\left(1+\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&(1-\\alpha S^{\\mathsf v,+}_{\\lambda})^{-1}\\alpha\\end{bmatrix}\\begin{bmatrix}\nS\\!L^{+}_{\\lambda}\n(1-\\alpha S^{\\mathsf v,+}_{\\lambda})^{-1}\\alpha(S\\!L^{-}_{\\lambda})^{*}& S\\!L^{+}_{\\lambda}\n\\\\(S\\!L^{-}_{\\lambda})^{*}& 0\n\\end{bmatrix}\\right)\\times\\\\\n&\\times\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&(1-\\alpha S^{\\mathsf v,+}_{\\lambda})^{-1}\\alpha\\end{bmatrix}\\,,\n\\end{align*}\nwhere\n$$\nR^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}R_{\\lambda\\pm i\\epsilon}\\,,\\qquadS\\!L^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}S\\!L_{\\lambda\\pm i\\epsilon}\\,,\\qquad \nS^{\\mathsf v,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\gamma_{0}S\\!L^{\\mathsf v}_{\\lambda\\pm i\\epsilon}\\,.\n$$\n\\subsection{\\label{Sec_dirichlet} Short-range potentials and Dirichlet boundary conditions.} \nHere we take \n$$\n\\mathfrak h_{2}= B^{3\/2}_{2,2}(\\Gamma)\\hookrightarrow\\mathfrak b_{2}=H^{1\/2}(\\Gamma)\n\\hookrightarrow\\mathfrak h_{2}^{\\circ}=L^{2}(\\Gamma)\\hookrightarrow \\mathfrak b_{2,2}=\\mathfrak b_{2}^{*}= H^{-1\/2}(\\Gamma)\\,,\n$$\n$$\n\\tau_{2}=\\gamma_{0}:H^{2}(\\mathbb{R}^{3})\\to B^{3\/2}_{2,2}(\\Gamma)\\,,\\qquad\nB_{0}=0\\,,\\qquad B_{2}=1\\,.\n$$\nFor any $z\\in\\mathbb{C}\\backslash(-\\infty,0]$, one has\n$$\nM_{z}^{\\mathsf B\n=\\begin{bmatrix}\n1 & 0\\\\\n0 & 0%\n\\end{bmatrix}-\n\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & 1%\n\\end{bmatrix}\n\\begin{bmatrix}\n\\langle x\\rangle^{-s}R_{z}\\langle x\\rangle^{-s} & \\langle x\\rangle^{-s}R_{z}\\gamma_{0}^{*}\\\\\n\\gamma_{0}R_{z}\\langle x\\rangle^{-s} & \\gamma_{0}R_{z}\\gamma_{0}^{*}%\n\\end{bmatrix}=\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}M_{z}^{\\mathsf v,d}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\\,,\n$$\n$$\nM_{z}^{\\mathsf v,d}:=%\n\\begin{bmatrix}\n1-\\mathsf v R_{z} & -\\mathsf v S\\!L_{z}\\\\\n-S\\!L_{\\bar z}^{*} & -S_{z}%\n\\end{bmatrix}\n\\,.\n$$\nBy the mapping properties provided in Sections \\ref{Sec_V} and \\ref{Sec_Layer}, by \\eqref{v-w} and \\eqref{R-w} with $w=-s$, one gets\n$$\nM_{z}^{\\mathsf v,d}\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{-1\/2}(\\Gamma),H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{1\/2}(\\Gamma))\\,.\n$$\nBy Lemma \\ref{coerc} with $\\mathsf v=0$, for any $z\\in Z^{\\circ}_{0,d}\\not=\\varnothing $, \n$$\n(M_{z}^{B_{0},B_{2}})^{-1}=\\Lambda_{z}^{B_{0},B_{2}}=\n(M_{z}^{d})^{-1}=\\Lambda_{z}^{\\!d}:=-S_{z}^{-1}\\in\\mathscr B(H^{1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\,.\n$$\nThus,\n$$\nZ_{B_{0},B_{2}}=Z_{d}:=\\{z\\in \\mathbb{C}\\backslash(-\\infty,0]: (M_{z}^{d})^{-1}\\in\\mathscr B(H^{1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\}\\supseteq Z^{\\circ}_{0,d}\\,.\n$$\nBy Lemma \\ref{coerc} again, for any $z\\in Z^{\\circ}_{\\mathsf v,d}\\not=\\varnothing $, \n$$\n(\\widehat M_{z}^{B_{0},B_{2}})^{-1}=(\\widehat \\Lambda_{z}^{B_{0},B_{2}})^{-1}=(\\widehat M_{z}^{\\mathsf v,d})^{-1}=\\widehat \\Lambda_{z}^{\\mathsf v,d}:=-(S^{\\mathsf v}_{z})^{-1}\\in\\mathscr B(H^{1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\,.\n$$\nThus,\n$$\n\\widehat Z_{\\mathsf B}=\\widehat Z_{\\mathsf v,d}:=\\{z\\in \\varrho(\\Delta+\\mathsf v): (\\widehat M_{z}^{\\mathsf v,d})^{-1}\\in\\mathscr B(H^{1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\}\\supseteq Z^{\\circ}_{\\mathsf v,d}\\,.\n$$\nHence,\n\\begin{align*}\n\\Lambda^{\\! \\mathsf B}_{z}=\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}(M_{z}^{\\mathsf v,d})^{-1}\n\\begin{bmatrix}\\langle x\\rangle^{-s}&0\\\\0&1\\end{bmatrix}\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & 1%\n\\end{bmatrix}=\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{\\Lambda}_{z}^{\\!\\mathsf v,d}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\\,,\n\\end{align*}\nwhere, by \\eqref{LB-new} and \\eqref{LB-new2},\n\\begin{align*}\n{\\Lambda}_{z}^{\\!\\mathsf v,d}:=&\\begin{bmatrix}\n\\Lambda_{z}^{\\!\\mathsf v}-\\Lambda_{z}^{\\!\\mathsf v}S\\!L_{z}(S^{\\mathsf v}_{z})^{-1}S\\!L_{\\bar z}^{*}\\Lambda_{z}^{\\!\\mathsf v}& -\\Lambda_{z}^{\\!\\mathsf v}S\\!L_{z}(S^{\\mathsf v}_{z})^{-1}\\\\\n-(S^{\\mathsf v}_{z})^{-1}S\\!L_{\\bar z}^{*}\\Lambda_{z}^{\\!\\mathsf v} & -(S^{\\mathsf v}_{z})^{-1}\n\\end{bmatrix}\\\\\n=&\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&-(S^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\n-S\\!L_{z}(S^{\\mathsf v}_{z})^{-1}S\\!L_{\\bar z}^{*}& S\\!L_{z}\n\\\\ S\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&-(S^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\end{align*}\nOne has \n\\begin{equation}\\label{Lbbv-dir}\n{\\Lambda}_{z}^{\\!\\mathsf v,d}\n\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)\\oplus H^{1\/2}(\\Gamma),H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{-1\/2}(\\Gamma))\\,.\n\\end{equation}\nBy Theorems \\ref{Th_Krein} and \\ref{Th-alt-res}, there follows that\n\\begin{align}\n&R_{z}^{\\mathsf v,d}=R_{z}+\n\\begin{bmatrix}R_{z}\\langle x\\rangle^{-s}&S\\!L_{z}\\end{bmatrix}\n\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{\\Lambda}_{z}^{\\!\\mathsf v,d}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\n\\begin{bmatrix}\\langle x\\rangle^{2s}\\mathsf v\\langle x\\rangle^{-s} R_{z}\\\\S\\!L^{*}_{\\bar z}\\end{bmatrix}\\label{Rv-dir-0}\n\\\\\n=&R_{z}+\n\\begin{bmatrix}R_{z}&S\\!L_{z}\\end{bmatrix}\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&-(S^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\n-S\\!L_{z}(S^{\\mathsf v}_{z})^{-1}S\\!L_{\\bar z}^{*}& S\\!L_{z}\n\\\\ S\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&-(S^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\begin{bmatrix}R_{z}\\\\S\\!L^{*}_{\\bar z}\\end{bmatrix}\\label{Rv-dir-1}\n\\\\\n=&R_{z}^{\\mathsf v}-S\\!L^{\\mathsf v}(S_{z}^{\\mathsf v})^{-1}{S\\!L_{\\bar z}^{\\mathsf v}}^{*}\n\\label{Rv-dir-2}\n\\end{align}\nis the resolvent of a self-adjoint operator $\\Delta^{\\!\\mathsf v,d}$; \\eqref{Rv-dir-0} holds for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,d})\\cap\\mathbb{C}\\backslash(-\\infty,0]$, both \n\\eqref{Rv-dir-1} and \n\\eqref{Rv-dir-2} hold for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,d})\\cap\\varrho(\\Delta+\\mathsf v)$.\n\\par\nBy Theorem \\ref{Th-add} and by $[\\gamma_{1}]u=-\\widehat\\rho_{\\mathsf B}u$ for any $u\\in\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,d})$,\n$$\n\\Delta^{\\!\\mathsf v,d}\\,u=\\Delta u+\\mathsf v u-([\\gamma_{1}]u)\\delta_{\\Gamma}\\,.\n$$\nBy \\eqref{Rv-alpha-2} and by the mapping properties of $S\\!L^{\\mathsf v}_{z}$, one has \n$$\n\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,d})\\subseteq H^{1}(\\mathbb{R}^{3})\\,.\n$$ \nMoreover, by Lemma \\ref{alt-abc}, \n$$\nu\\in\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,d})\\quad\\Longrightarrow\\quad\\gamma_{0}u=0\\,.\n$$\nTherefore, $\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,d})\\subseteq H_{0}^{1}(\\Omega_{\\rm in})\\oplus H_{0}^{1}(\\Omega_{\\rm ex})$.\nSince $\\widehat Z_{\\mathsf v,\\alpha}$ contains a positive half-line, $\\Delta^{\\!\\mathsf v,d}$ is bounded from above and hypothesis (H4.1) holds. The scattering couple $(\\Delta^{\\!\\mathsf v,d},\\Delta)$ is asymptotically complete and the corresponding scattering matrix is given by \n$$\n{\\mathcal S}_{\\lambda}^{\\mathsf v,d}=1-2\\pi iL_{\\lambda}\\Lambda^{\\mathsf v,d,+}_{\\lambda}L_{\\lambda}^{*}\\,,\\quad \\lambda\\in(-\\infty,0]\\backslash(\\sigma^{-}_{p}(\\Delta+\\mathsf v)\\cup \\sigma^{-}_{p}(\\Delta^{\\!\\mathsf v,d}))\\,,\n$$\nwhere $L_{\\lambda}$ is given in Corollary \\ref{Llambda} and $\\Lambda^{\\mathsf v,d,+}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\Lambda^{\\mathsf v,d}_{\\lambda+i\\epsilon}$. This latter limit exists by Lemma \\ref{rmH7}; in particular, by \\eqref{LBpm2}, \n\\begin{align*}\n\\Lambda^{\\mathsf v,d,+}=&\\left(1+\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&-(S^{\\mathsf v,+}_{\\lambda})^{-1}\\end{bmatrix}\\begin{bmatrix}\n-S\\!L^{+}_{\\lambda}\n(S^{\\mathsf v,+}_{\\lambda})^{-1}(S\\!L^{-}_{\\lambda})^{*}& S\\!L^{+}_{\\lambda}\n\\\\(S\\!L^{-}_{\\lambda})^{*}& 0\n\\end{bmatrix}\\right)\\times\\\\\n&\\times\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&-(S^{\\mathsf v,+}_{\\lambda})^{-1}\\end{bmatrix}\\,,\n\\end{align*}\nwhere\n$$\nR^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}R_{\\lambda\\pm i\\epsilon}\\,,\\qquadS\\!L^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}S\\!L_{\\lambda\\pm i\\epsilon}\\,,\\qquad \nS^{\\mathsf v,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\gamma_{0}S\\!L^{\\mathsf v}_{\\lambda\\pm i\\epsilon}\\,.\n$$\n\\subsection{\\label{Sec_neumann} Short-range potentials and Neumann boundary conditions.} \nHere we take \n$$\n\\mathfrak h_{2}=\\mathfrak b_{2}^{*}=\\mathfrak b_{2,2}= H^{1\/2}(\\Gamma)\\hookrightarrow\\mathfrak h_{2}^{\\circ}=L^{2}(\\Gamma)\\hookrightarrow\\mathfrak b_{2}=\\mathfrak h^{*}_{2}=\\mathfrak b_{2,2}^{*}=H^{-1\/2}(\\Gamma)\n\\,,\n$$\n$$\n\\tau_{2}=\\gamma_{1}:H^{2}(\\mathbb{R}^{3})\\to H^{1\/2}(\\Gamma)\\,,\\qquad\nB_{0}=0\\,,\\qquad B_{2}=1\\,.\n$$\nFor any $z\\in\\mathbb{C}\\backslash(-\\infty,0]$, one has\n$$\nM_{z}^{\\mathsf B\n=\\begin{bmatrix}\n1 & 0\\\\\n0 & 0%\n\\end{bmatrix}-\n\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & 1%\n\\end{bmatrix}\n\\begin{bmatrix}\n\\langle x\\rangle^{-s}R_{z}\\langle x\\rangle^{-s} & \\langle x\\rangle^{-s}R_{z}\\gamma_{1}^{*}\\\\\n\\gamma_{1}R_{z}\\langle x\\rangle^{-s} & \\gamma_{1}R_{z}\\gamma_{0}^{*}%\n\\end{bmatrix}\n=\\begin{bmatrix}\\langle x\\rangle^{s} & 0\\\\\n0 & 1\n\\end{bmatrix}M_{z}^{\\mathsf v,n}\n\\begin{bmatrix}\\langle x\\rangle^{s} & 0\\\\\n0 & 1%\n\\end{bmatrix}\\,,\n$$\n$$\nM_{z}^{\\mathsf v,n}:=%\n\\begin{bmatrix}\n1-\\mathsf v R_{z} & -\\mathsf v D\\!L_{z}\\\\\n-D\\!L_{\\bar z}^{*} & -D_{z}%\n\\end{bmatrix}\n\\,.\n$$\nBy the mapping properties provided in Sections \\ref{Sec_V} and \\ref{Sec_Layer}, by \\eqref{v-w} and \\eqref{R-w} with $w=-s$, one gets\n$$\nM_{z}^{\\mathsf v,n}\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{1\/2}(\\Gamma),\nH_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{-1\/2}(\\Gamma))\\,.\n$$\nBy Lemma \\ref{coerc} with $\\mathsf v=0$, for any $z\\in Z^{\\circ}_{0,n}\\not=\\varnothing $, \n$$\n(M_{z}^{B_{0},B_{2}})^{-1}=\\Lambda_{z}^{B_{0},B_{2}}=\n(M_{z}^{n})^{-1}=\\Lambda_{z}^{\\!n}:=-D_{z}^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma))\\,.\n$$\nThus,\n$$\nZ_{B_{0},B_{2}}=Z_{n}:=\\{z\\in \\mathbb{C}\\backslash(-\\infty,0]: (M_{z}^{n})^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma))\\}\\supseteq Z^{\\circ}_{0,n}\\,.\n$$\nBy Lemma \\ref{coerc} again, for any $z\\in Z^{\\circ}_{\\mathsf v,n}\\not=\\varnothing $, \n$$\n(\\widehat M_{z}^{B_{0},B_{2}})^{-1}=(\\widehat \\Lambda_{z}^{B_{0},B_{2}})^{-1}=(\\widehat M_{z}^{\\mathsf v,n})^{-1}=\\widehat \\Lambda_{z}^{\\mathsf v,n}:=-(D^{\\mathsf v}_{z})^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma))\\,.\n$$\nThus,\n$$\n\\widehat Z_{\\mathsf B}=\\widehat Z_{n}:=\\{z\\in \\varrho(\\Delta+\\mathsf v): (\\widehat M_{z}^{\\mathsf v,n})^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{1\/2}(\\Gamma))\\}\\supseteq Z^{\\circ}_{\\mathsf v,n}\\,.\n$$\nHence, \n\\begin{align*}\n\\Lambda^{\\! \\mathsf B}_{z}=\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}(M_{z}^{\\mathsf v,n})^{-1}\n\\begin{bmatrix}\\langle x\\rangle^{-s}&0\\\\0&1\\end{bmatrix}\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & 1%\n\\end{bmatrix}=\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{\\Lambda}_{z}^{\\!\\mathsf v,n}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\\,,\n\\end{align*}\nwhere, by \\eqref{LB-new} and by \\eqref{LB-new2}\n\\begin{align*}\n{\\Lambda}_{z}^{\\!\\mathsf v,n}:=&\\begin{bmatrix}\n\\Lambda_{z}^{\\!\\mathsf v}-\\Lambda_{z}^{\\!\\mathsf v}D\\!L_{z}(D^{\\mathsf v}_{z})^{-1}D\\!L_{\\bar z}^{*}\\Lambda_{z}^{\\!\\mathsf v}& -\\Lambda_{z}^{\\!\\mathsf v}D\\!L_{z}(D^{\\mathsf v}_{z})^{-1}\\\\\n-(D^{\\mathsf v}_{z})^{-1}D\\!L_{\\bar z}^{*}\\Lambda_{z}^{\\!\\mathsf v} & -(D^{\\mathsf v}_{z})^{-1}\n\\end{bmatrix}\\\\\n=&\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&-(D^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\n-D\\!L_{z}(D^{\\mathsf v}_{z})^{-1}D\\!L_{\\bar z}^{*}& D\\!L_{z}\n\\\\ D\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&-(D^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\end{align*}\nOne has \n\\begin{equation}\\label{Lbbv-neu}\n{\\Lambda}_{z}^{\\!\\mathsf v,n}\n\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)\\oplus H^{-1\/2}(\\Gamma),H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{1\/2}(\\Gamma))\\,.\n\\end{equation}\nBy Theorems \\ref{Th_Krein} and \\ref{Th-alt-res}, there follows that \n\\begin{align}\n&R_{z}^{\\mathsf v,n}\n=R_{z}+\n\\begin{bmatrix}R_{z}\\langle x\\rangle^{-s}&D\\!L_{z}\\end{bmatrix}\n\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{\\Lambda}_{z}^{\\!\\mathsf v,n}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\\begin{bmatrix}\\langle x\\rangle^{2s}\\mathsf v \\langle x\\rangle^{-s}R_{z}\\\\ D\\!L^{*}_{\\bar z}\\end{bmatrix}\\label{Rv-neu-0}\\\\ \n=&R_{z}+\n\\begin{bmatrix}R_{z}&D\\!L_{z}\\end{bmatrix}\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\!\\!\\!-(D^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\n-D\\!L_{z}(D^{\\mathsf v}_{z})^{-1}D\\!L_{\\bar z}^{*}& D\\!L_{z}\n\\\\ D\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\!\\!\\!-(D^{\\mathsf v}_{z})^{-1}\\end{bmatrix}\n\\begin{bmatrix}R_{z}\\\\ D\\!L^{*}_{\\bar z}\\end{bmatrix}\\label{Rv-neu-1}\\\\ \n=&R_{z}^{\\mathsf v}-D\\!L_{z}^{\\mathsf v}(D_{z}^{\\mathsf v})^{-1}{D\\!L^{\\mathsf v}_{\\bar z}}^{*}\n\\label{Rv-neu-2}\n\\end{align}\nis the resolvent of a self-adjoint operator $\\Delta^{\\!\\mathsf v,n}$; \\eqref{Rv-neu-0} holds for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,n})\\cap\\mathbb{C}\\backslash(-\\infty,0]$, both \\eqref{Rv-neu-1} and \\eqref{Rv-neu-2} hold for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,n})\\cap\\varrho(\\Delta+\\mathsf v)$.\n\\par \nBy Theorem \\ref{Th-add} and by $[\\gamma_{0}]u=\\widehat\\rho_{\\mathsf B}u$ for any $u\\in\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,n})$,\n$$\n\\Delta^{\\!\\mathsf v,n}\\,u=\\Delta u+\\mathsf v u+([\\gamma_{0}]u)\\delta'_{\\Gamma}\\,.\n$$\nBy \\eqref{Rv-alpha-2} and by the mapping properties of $D\\!L^{\\mathsf v}_{z}$, one has \n$$\n\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,n})\\subseteq H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)\\,.\n$$ \nMoreover, by Lemma \\ref{alt-abc}, \n$$\nu\\in\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,n})\\quad\\Longrightarrow\\quad\\gamma_{1}u=0\\,.\n$$\nSince $\\widehat Z_{\\mathsf v,n}$ contains a positive half-line, $\\Delta^{\\!\\mathsf v,n}$ is bounded from above and hypothesis (H4.1) holds. The scattering couple $(\\Delta^{\\!\\mathsf v,n},\\Delta)$ is asymptotically complete and the corresponding scattering matrix is given by \n$$\n{\\mathcal S}_{\\lambda}^{\\mathsf v,n}=1-2\\pi iL_{\\lambda}\\Lambda^{\\mathsf v,n,+}_{\\lambda}L_{\\lambda}^{*}\\,,\\quad \\lambda\\in(-\\infty,0]\\backslash(\\sigma^{-}_{p}(\\Delta+\\mathsf v)\\cup \\sigma^{-}_{p}(\\Delta^{\\!\\mathsf v,n}))\\,,\n$$\nwhere $L_{\\lambda}$ is given in Corollary \\ref{Llambda} and $\\Lambda^{\\mathsf v,n,+}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\Lambda^{\\mathsf v,n}_{\\lambda+i\\epsilon}$. This latter limit exists by Lemma \\ref{rmH7}; in particular, by \\eqref{LBpm2}, \n\\begin{align*}\n\\Lambda^{\\mathsf v,n,+}=&\\left(1+\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&-(D^{\\mathsf v,+}_{\\lambda})^{-1}\\end{bmatrix}\\begin{bmatrix}\n-D\\!L^{+}_{\\lambda}\n(D^{\\mathsf v,+}_{\\lambda})^{-1}(D\\!L^{-}_{\\lambda})^{*}& D\\!L^{+}_{\\lambda}\n\\\\(D\\!L^{-}_{\\lambda})^{*}& 0\n\\end{bmatrix}\\right)\\times\\\\\n&\\times\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&-(D^{\\mathsf v,+}_{\\lambda})^{-1}\\end{bmatrix}\\,,\n\\end{align*}\nwhere\n$$\nR^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}R_{\\lambda\\pm i\\epsilon}\\,,\\qquadD\\!L^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}D\\!L_{\\lambda\\pm i\\epsilon}\\,,\\qquad \nD^{\\mathsf v,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\gamma_{1}D\\!L^{\\mathsf v}_{\\lambda\\pm i\\epsilon}\\,.\n$$\n\\subsection{\\label{Sec_delta'} Short-range potentials and semi-transparent boundary conditions\nof $\\delta'_{\\Gamma }$-type} \nHere we take \n$$\n\\mathfrak h_{2}=\\mathfrak b_{2}^{*}=\\mathfrak b_{2,2}= H^{1\/2}(\\Gamma)\\hookrightarrow\\mathfrak h_{2}^{\\circ}=L^{2}(\\Gamma)\\hookrightarrow \\mathfrak b_{2}=\\mathfrak h^{*}_{2}=\\mathfrak b_{2,2}^{*}=H^{-1\/2}(\\Gamma)\n\\,,\n$$\n$$\n\\tau_{2}=\\gamma_{1}:H^{2}(\\mathbb{R}^{3})\\to H^{1\/2}(\\Gamma)\\,,\\qquad\nB_{0}=\\theta\\,,\\qquad B_{2}=1\\,,\n$$\nwhere \n$$ \n\\theta\\in\\mathscr B(H^{s_{\\circ}}(\\Gamma),H^{-s_{\\circ}}(\\Gamma))\\,,\\quad \n02$, fulfills our hypothesis. \nLet us also remark that $\\mathscr B(H^{s_{\\circ}}(\\Gamma),H^{-s_{\\circ}}(\\Gamma))\\subseteq\n\\mathscr B(H^{1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))=\\mathscr B(\\mathfrak b_{2}^{*},\\mathfrak b_{2,2}^{*})$.\\par\nFor any $z\\in\\mathbb{C}\\backslash(-\\infty,0]$, one has\n$$\nM_{z}^{\\mathsf B\n=\\begin{bmatrix}\n 1 & 0\\\\\n0 & \\theta%\n\\end{bmatrix}\n-\n\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & 1%\n\\end{bmatrix}\n\\begin{bmatrix}\n\\langle x\\rangle^{-s}R_{z}\\langle x\\rangle^{-s} & \\langle x\\rangle^{-s}R_{z}\\gamma_{1}^{*}\\\\\n\\gamma_{1}R_{z}\\langle x\\rangle^{-s} & \\gamma_{1}R_{z}\\gamma_{1}^{*}%\n\\end{bmatrix}=\n\\begin{bmatrix}\\langle x\\rangle^{s} & 0\\\\\n0 & 1\\end{bmatrix} \nM_{z}^{\\mathsf v,\\theta}\n\\begin{bmatrix}\\langle x\\rangle^{-s} & 0\\\\\n0 & 1\\end{bmatrix}\\,,\n$$\n$$\nM_{z}^{\\mathsf v,\\theta}:=%\n\\begin{bmatrix}\n1-\\mathsf v R_{z} & -\\mathsf v D\\!L_{z}\\\\\n-D\\!L_{\\bar z}^{*} & \\theta-D_{z}\n\\end{bmatrix}\\,.%\n$$\nBy the mapping properties provided in Sections \\ref{Sec_V} and \\ref{Sec_Layer}, by \\eqref{v-w} and \\eqref{R-w} with $w=-s$, one gets\n$$\nM_{z}^{\\mathsf v,\\theta}\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{1\/2}(\\Gamma),H_{-s}^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{-1\/2}(\\Gamma))\\,.\n$$\n\\begin{lemma}\\label{LLt} \n$$\n\\forall z\\in \\widehat Z^{\\circ}_{\\mathsf v,n}:=Z^{\\circ}_{\\mathsf v,n}\\cap\\mathbb{C}\\backslash\\mathbb{R}\\,,\\qquad (1-\\theta (D^{\\mathsf v}_{z})^{-1})^{-1}\\in \\mathscr B( H^{-1\/2}( \\Gamma)) \\,.\n$$\n\\end{lemma}\n\\begin{proof}\nWe follow the same the arguments as in the proof of \\cite[Lemma 5.4]{JMPA}. Since, by the compact embedding $H^{-s_{\\circ}}(\\Gamma)\\hookrightarrow H^{-1\/2}(\\Gamma)$, $\\theta( D_{z}^{\\mathsf v})^{-1}\\in{\\mathscr B}(H^{-1\/2}(\\Gamma))$ is compact, \nby the Fredholm alternative, $1-\\theta( D_{z}^{\\mathsf v}) ^{-1}$ has a bounded inverse if and only if it has trivial kernel. Let $\\varphi\\in H^{-1\/2}(\\Gamma)$ be such that $D_{z}^{\\mathsf v}\\varphi=\\theta\\varphi$; using the self-adjointness of $\\theta$, we get%\n\\[\n( D_{z}^{\\mathsf v}-D_{\\bar z}^{\\mathsf v}) \\varphi=0\\,.\n\\]\nBy the resolvent identity,\n\\[\n\\text{Im}(z)\\gamma_{1}R_{\\bar z}^{\\mathsf v}R_{z}^{\\mathsf v}%\n\\gamma_{1}^{\\ast}\\varphi=0\\,.\n\\]\nThis gives \n\\begin{equation}\n\\|R_{z}^{\\mathsf v}\\gamma_{1}^{\\ast}\\varphi\\| _{L^{2}(\\mathbb{R}^{3})}=0\\,.\n\\end{equation}\nSince $(R_{z}^{\\mathsf v}\\gamma_{1}^{\\ast})^{\\ast}=\\gamma_{1}R_{\\bar\n{z}}^{\\mathsf v}\\in{\\mathscr B}( L^{2}( \\mathbb{R}^{3}),H^{1\/2}(\\Gamma)) $ is surjective, then\n$R_{z}^{\\mathsf v}\\gamma_{1}^{\\ast}\\in{\\mathscr B}( H^{-1\/2}( \\Gamma) ,L^{2}( \\mathbb{R}^{3})) $ has closed\nrange by the closed range theorem and, by \\cite[Theorem 5.2, p. 231]{Kato},\n\\[\n\\|R_{z}^{\\mathsf v}\\gamma_{1}^{\\ast}\\varphi\\|_{L^{2}(\n\\mathbb{R}^{3}) }\\gtrsim\\|\\varphi\\|_{H^{-1\/2}(\\Gamma) }\\,.\n\\]\nThus $\\text{\\rm ker}(1-\\theta (D_{z}^{\\mathsf v})^{-1})=\\{0\\}$ and the proof is done.\n\\end{proof}\nAccording to Lemma \\ref{LLt} with $\\mathsf v=0$, for any $z\\in \\widehat Z^{\\circ}_{0,n}\\not=\\varnothing $,\n$$\n(M_{z}^{B_{0},B_{2}})^{-1}=(M_{z}^{\\theta})^{-1}=\\Lambda_{z}^{\\!\\theta}:=(\\theta-D_{z})^{-1}=\n-D_{z}^{-1}(1-\\theta D_{z}^{-1})^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\,.\n$$\nThus\n$$\nZ_{B_{0},B_{2}}=Z_{\\theta}:=\\{z\\in \\mathbb{C}\\backslash(-\\infty,0]:(M_{z}^{\\theta})^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\}\\supseteq\\widehat Z^{\\circ}_{0,n}\\,.\n$$\nAccording to Lemma \\ref{LLt} again, for any $z\\in \\widehat Z^{\\circ}_{\\mathsf v,n}\\not=\\varnothing $,\n$$\n(\\widehat M_{z}^{B_{0},B_{2}})^{-1}=(\\widehat M_{z}^{\\mathsf v,\\theta})^{-1}=\\widehat \\Lambda_{z}^{\\mathsf v,\\theta}:=(\\theta-D_{z}^{\\mathsf v})^{-1}=\n-(D^{\\mathsf v}_{z})^{-1}(1-\\theta (D^{\\mathsf v}_{z})^{-1})^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\,.\n$$\nThus\n$$\n\\widehat Z_{\\mathsf B}=\\widehat Z_{\\mathsf v,\\theta}:=\\{z\\in \\varrho(\\Delta+\\mathsf v):(\\widehat M_{z}^{\\mathsf v,\\theta})^{-1}\\in\\mathscr B(H^{-1\/2}(\\Gamma),H^{-1\/2}(\\Gamma))\\}\\supseteq\\widehat Z^{\\circ}_{\\mathsf v,n}\\,.\n$$\nHence, \n$$\n\\Lambda_{z}^{\\mathsf B}=\\begin{bmatrix}\\langle x\\rangle^{s} & 0\\\\\n0 & 1\\end{bmatrix} \n(M_{z}^{\\mathsf v,\\theta})^{-1}\n\\begin{bmatrix}\\langle x\\rangle^{-s} & 0\\\\\n0 & 1\\end{bmatrix}\\begin{bmatrix}\n\\langle x\\rangle^{2s}\\mathsf v & 0\\\\\n0 & 1\n\\end{bmatrix}=\\begin{bmatrix}\\langle x\\rangle^{s} & 0\\\\\n0 & 1\\end{bmatrix} \n\\Lambda_{z}^{\\!\\mathsf v,\\theta}\n\\begin{bmatrix}\\langle x\\rangle^{s} & 0\\\\\n0 & 1\\end{bmatrix}\\,,\n$$\nwhere, by \\eqref{LB-new} and by \\eqref{LB-new2},\n\\begin{align*}\n\\Lambda_{z}^{\\!\\mathsf v,\\theta}:=&\\begin{bmatrix}\n\\Lambda_{z}^{\\!\\mathsf v,\\theta}\\Lambda^{\\!\\mathsf v}_{z}+\\Lambda^{\\!\\mathsf v}_{z} D\\!L_{z}\\widehat\\Lambda^{\\mathsf v,\\theta}_{z}D\\!L_{\\bar z}^{*}\\Lambda^{\\!\\mathsf v}_{z}\n& \\Lambda^{\\mathsf v}_{z} D\\!L_{z}\\widehat\\Lambda^{\\mathsf v,\\theta}_{z}\\\\\n\\widehat\\Lambda^{\\mathsf v,\\theta}_{z}D\\!L_{\\bar z}^{*}\\Lambda^{\\!\\mathsf v}_{z}& \\widehat\\Lambda^{\\mathsf v,\\theta}_{z}\\end{bmatrix}\\\\\n=&\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\theta}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\nD\\!L_{z}\\widehat\\Lambda_{z}^{\\mathsf v,\\theta}D\\!L_{\\bar z}^{*}& D\\!L_{z}\n\\\\ D\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\theta}\\end{bmatrix}\n\\,.\n\\end{align*}\nOne has \n\\begin{equation}\\label{Ltbbteta}\n{\\Lambda}_{z}^{\\!\\mathsf v,\\theta}\n\\in{\\mathscr B}(H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)\\oplus H^{-1\/2}(\\Gamma),H_{-s}^{1}( \\mathbb{R}^{3}\\backslash\\Gamma)^{*}\\oplus H^{1\/2}(\\Gamma))\\,.\n\\end{equation}\nBy Theorems \\ref{Th_Krein} and \\ref{Th-alt-res}, there follows that\n\\begin{align}\nR_{z}^{\\mathsf v,\\theta}\n=&R_{z}+\\begin{bmatrix}R_{z}\\langle x\\rangle^{-s}&D\\!L_{z}\\end{bmatrix}\n\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}{\\Lambda}_{z}^{\\!\\mathsf v,\\theta}\\begin{bmatrix}\\langle x\\rangle^{s}&0\\\\0&1\\end{bmatrix}\n\\begin{bmatrix}\\langle x\\rangle^{2s}\\mathsf v \\langle x\\rangle^{-s}R_{z}\\\\D\\!L^{*}_{\\bar z}\\end{bmatrix}\n\\label{Rv-teta-0}\\\\\n=&R_{z}+\n\\begin{bmatrix}R_{z}&D\\!L_{z}\\end{bmatrix}\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\theta}\\end{bmatrix}\n\\left(1+\\begin{bmatrix}\nD\\!L_{z}\\widehat\\Lambda_{z}^{\\mathsf v,\\theta}D\\!L_{\\bar z}^{*}& D\\!L_{z}\n\\\\ D\\!L_{\\bar z}^{*} & 0\n\\end{bmatrix}\\right)\n\\begin{bmatrix}\\Lambda^{\\mathsf v}_{z}&0\\\\0&\\widehat\\Lambda_{z}^{\\mathsf v,\\theta}\\end{bmatrix}\n\\begin{bmatrix}R_{z}\\\\D\\!L^{*}_{\\bar z}\\end{bmatrix}\n\\label{Rv-teta-1}\\\\ \n=&R_{z}^{\\mathsf v}+D\\!L_{z}^{\\mathsf v}{\\widehat\\Lambda}_{z}^{\\mathsf v,\\theta}{D\\!L_{\\bar z}^{\\mathsf v}}^{*}\\,.\n\\label{Rv-teta-2}\n\\end{align}\nis the resolvent of a self-adjoint operator $\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta}$; \\eqref{Rv-teta-0} holds for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta})\\cap\\mathbb{C}\\backslash(-\\infty,0]$, both \\eqref{Rv-teta-1} and \\eqref{Rv-teta-2} hold for any \n$z\\in \\varrho(\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta})\\cap\\varrho(\\Delta+\\mathsf v)$.\n\\par\nBy Theorem \\ref{Th-add},\n$$\n\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta}u=\\Delta u+\\mathsf v u+(\\theta\\gamma_{1}u)\\delta'_{\\Gamma}\\,.\n$$\nBy \\eqref{Rv-alpha-2} and by the mapping properties of $D\\!L^{\\mathsf v}_{z}$, one has \n$$\n\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta})\\subseteq H^{1}(\\mathbb{R}^{3}\\backslash\\Gamma)\\,.\n$$ \nMoreover, by $R^{\\mathsf v}_{z}u\\in H^{2}(\\mathbb{R}^{3})$, so that $[\\gamma_{1}]R^{\\mathsf v}_{z}u=0$, and by \\eqref{jumpv1}, one gets $[\\gamma_{0}]R^{\\mathsf v,\\theta}_{z}u={\\widehat\\Lambda}_{z}^{\\mathsf v,\\theta}{D\\!L_{\\bar z}^{\\mathsf v}}^{*}u=\\widehat\\rho_{\\mathsf B}(R^{\\mathsf v,\\theta}_{z}u)$. Hence, by Lemma \\ref{alt-abc}, \n$$\nu\\in\\text{\\rm dom}(\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta})\\quad\\Longrightarrow\\quad\\gamma_{1}u=\\theta[\\gamma_{0}]u\\,.\n$$\nProceeding as in \\cite[Subsection 5.5]{JMPA} (see the proof of Theorem 5.15 there), $\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta}$ is bounded from above and so hypothesis (H4.1) holds. The scattering couple $(\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta},\\Delta)$ is asymptotically complete and the corresponding scattering matrix is given by \n$$\n{\\mathcal S}_{\\lambda}^{\\mathsf v,\\theta}=1-2\\pi iL_{\\lambda}\\Lambda^{\\mathsf v,\\theta,+}_{\\lambda}L_{\\lambda}^{*}\\,,\\quad \\lambda\\in(-\\infty,0]\\backslash(\\sigma^{-}_{p}(\\Delta+\\mathsf v)\\cup \\sigma^{-}_{p}(\\Delta^{\\!\\mathsf v,\\delta'\\!,\\theta}))\\,,\n$$\nwhere $L_{\\lambda}$ is given in Corollary \\ref{Llambda} and $\\Lambda^{\\mathsf v,\\theta,+}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\Lambda^{\\mathsf v,\\theta}_{\\lambda+i\\epsilon}$. This latter limit exists by Lemma \\ref{rmH7}; in particular, by \\eqref{LBpm2}, \n\\begin{align*}\n\\Lambda^{\\mathsf v,\\theta,+}=&\\left(1+\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&(\\theta-D^{\\mathsf v,+}_{\\lambda})^{-1}\\end{bmatrix}\\begin{bmatrix}\nD\\!L^{+}_{\\lambda}\n(\\theta-D^{\\mathsf v,+}_{\\lambda})^{-1}\\alpha(D\\!L^{-}_{\\lambda})^{*}& D\\!L^{+}_{\\lambda}\n\\\\(D\\!L^{-}_{\\lambda})^{*}& 0\n\\end{bmatrix}\\right)\\times\\\\\n&\\times\\begin{bmatrix}\n(1-\\mathsf v R^{+}_{\\lambda})^{-1}\\mathsf v&0\\\\\n0&(\\theta-D^{\\mathsf v,+}_{\\lambda})^{-1}\\alpha\\end{bmatrix}\\,,\n\\end{align*}\nwhere\n$$\nR^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}R_{\\lambda\\pm i\\epsilon}\\,,\\qquadD\\!L^{\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}S\\!L_{\\lambda\\pm i\\epsilon}\\,,\\qquad \nD^{\\mathsf v,\\pm}_{\\lambda}:=\\lim_{\\epsilon\\searrow 0}\\gamma_{0}D\\!L^{\\mathsf v}_{\\lambda\\pm i\\epsilon}\\,.\n$$\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nThere is a commonly accepted idea that the sunspot activity is produced\nby the large-scale toroidal magnetic field which is generated inside\nthe convection zone by means of the differential rotation \\citep{P55}.\nThe theory explains the 11-year solar cycle as a result of the large-scale\ndynamo operating in the solar interior, where, in addition to the\nmagnetic fields generated by the differential rotation, the helical\nconvective motions transforms the energy of the toroidal magnetic\nfields to poloidal. The effect of meridional circulation on the large-scale\ndynamo is not well understood. It is the essential part of the flux-transport\ndynamo model scenario \\citep{choud95,cd99} to explain the equatorward\ndrift of the toroidal magnetic field in the solar cycle. Here, it\nassumed that the toroidal field at the bottom of the convection zone\nforms sunspot activity. Feasibility of this idea can be questioned\nboth the observational and theoretical arguments \\citep{b05}. The\ndistributed dynamo models can be constructed with \\citep{2002AA...390..673B,2008AA...483..949J,2014AA563A18P}\nand without \\citep{moss00M,pip13M} effect of meridional transport\nof the large-scale magnetic field.\n\nRecent results of helioseismology reveal the double-cell meridional\ncirculation structure \\citep{Zhao13m,2017ApJ845.2B}. It demolishes\nthe previously accepted scenario of the flux-transport models \\citep{2014ApJ782.93H,2016MNRAS.456.2654W,2016ApJ832.9H}.\nContrary, \\citet{PK13} showed the distributed dynamo models can reproduce\nobservations with regards to the subsurface rotational shear layer\nand the double-cell meridional circulation. In their model, the double-cell\nmeridional circulation was modeled in following to results of helioseismology\nof \\citet{Zhao13m}. The effect of the multi-cell meridional circulation\non the global dynamo was also studied in the direct numerical simulations\n\\citep{kap2012,2016ApJ819.104G,2018AA609A..51W}. There were no attempts\nto construct the non-kinematic mean-field dynamo models with regards\nto the multi-cell meridional circulation.\n\nThe standard mean-field models of the solar differential rotation\npredict a one-cell meridional circulation per hemisphere. This contradicts\nto the helioseismology inversions and results of direct numerical\nsimulations. In the mean-field theory framework, the differential\nrotation of the Sun is explained as a result of the angular momentum\ntransport by the helical convective motions. Similarly to a contribution\nof the $\\alpha$ effect in the mean-electromotive force, i.e., \n\\[\n\\mathbf{\\mathcal{E}}=\\left\\langle \\mathbf{u}\\times\\mathbf{b}\\right\\rangle =\\hat{\\alpha}\\circ\\left\\langle \\mathbf{B}\\right\\rangle +\\dots,\n\\]\nwhere $\\mathbf{u}$ is the turbulent velocity $\\mathbf{u}$, and $\\mathbf{b}$\nis the turbulent magnetic field, the $\\Lambda$-effect, (e.g., \\citealp{1989drsc.book.....R})\nappears as the non-dissipative part of turbulent stresses\\textbf{\n\\[\n\\hat{T}_{ij}=\\left\\langle u_{i}u_{j}\\right\\rangle =\\Lambda_{ijk}\\Omega_{k}+\\dots\n\\]\n}where $\\boldsymbol{\\Omega}$ is the angular velocity. The structure\nof the meridional circulation is determined by directions of the non-diffusive\nangular momentum transport due to the $\\Lambda$ effect \\citep{2017ApJ835.9B}.\nIn particular, the vertical structure of the meridional circulation\ndepends on the sign of the radial effect. It was found that the double-cell\nmeridional circulation can be explained if the of $\\Lambda$-effect\nchanges sign in the depth of the convection zone. \\citet{2018ApJ854.67P}\nshowed that this effect can result from the radial inhomogeneity of\nthe convective turnover timescale. It was demonstrated that if this\neffect is taken into account then the solar-like differential rotation\nand the double-cell meridional circulation are both reproduced by\nthe mean-field model .\n\nIn this paper, we apply the meridional circulation profile, which\nis calculated from the solution of the angular momentum balance to\nthe nonkinematic dynamo models. Previously, the similar approach was\napplied by \\citet{1992AA...265..328B} and \\citet{2006ApJ...647..662R}\nin the distributed and the flux-transport models with one meridional\ncirculation cell as the basic stage in the non-magnetic case.\n\nOur main goal is to study how the double-cell meridional circulation\naffects the nonlinear dynamo generation of the large-scale magnetic\nfield. The magnetic feedback on the global flow can result in numerous\nphysical phenomena such as the torsional oscillations \\citep{1982SoPh...75..161L,2011JPhCS271a2074H},\nthe long-term variability of the magnetic activity \\citep{sok1994AA,2014JGRA119.6027F}\netc. The properties of the nonlinear evolution depend on the dynamo\ngoverning parameters such as amplitude of turbulent generation of\nthe magnetic field by the $\\alpha$ effect, as well as the other nonlinear\nprocesses involved in the dynamo, i.e., the dynamo quenching by the\nmagnetic buoyancy effect \\citep{kp93,tob98} and the magnetic helicity\nconservation \\citep{kleruz82}. We study if the long-term variation\nof magnetic activity can result from the increasing level of turbulent\ngeneration of magnetic field by the $\\alpha$ effect. The increasing\nof the $\\alpha$ effect results to an increase of the magnetic helicity\nproduction. This affects the large-scale magnetic field generation\nby means of the magnetic helicity conservation. Hence, the magnetic\nhelicity balance has to be taken into account. \n\n{It is hardly possible to consider in full all the goals within\none paper. From our point of view, the most important tasks includes:\nconstruction of the solar-type dynamo model with the multi-cell meridional\ncirculation and studying the principal nonlinear dynamo effects. The\nlatter includes the magnetic helicity conservation and the nonkinematic\neffects due to the magnetic feedback on the large-scale flow. Accordingly,\nthe paper is organized as follows.} Next Section describes the hydrodynamic,\nthermodynamic and magnetohydrodynamic parts of the model. Then, I\npresent an attempt to construct the solar-type dynamo model and discuss\nthe effect of the turbulent pumping on the properties of the dynamo\nsolution. {The next subsections consider result for the principal\nnonlinear dynamo effects. They deals with the global flows variations,\nthe Grand activity cycles and the magnetic cycle variations of the\nthermodynamic parameters in the model.} The paper is concluded with\na discussion of the main results using results of other theoretical\nstudies and results of observations.\n\n\\section{Basic equations.}\n\n\\subsection{The angular momentum balance\\label{subsec:am}}\n\nWe consider the evolution of the axisymmetric large-scale flow, which\nis decomposed into poloidal and toroidal components: $\\mathbf{\\overline{U}}=\\mathbf{\\overline{U}}^{m}+r\\sin\\theta\\Omega\\hat{\\mathbf{\\boldsymbol{\\phi}}}$,\nwhere $\\boldsymbol{\\hat{\\phi}}$ is the unit vector in the azimuthal\ndirection. The mean flow satisfies the stationary continuity equation,\n\\begin{equation}\n\\boldsymbol{\\nabla}\\cdot\\overline{\\rho}\\mathbf{\\overline{U}}=0,\\label{eq:cont}\n\\end{equation}\nDistribution of the angular velocity inside convection zone is determined\nby conservation of the angular momentum \\citep{1989drsc.book.....R}:\n\\begin{eqnarray}\n\\frac{\\partial}{\\partial t}\\overline{\\rho}r^{2}\\sin^{2}\\theta\\Omega & = & -\\boldsymbol{\\nabla\\cdot}\\left(r\\sin\\theta\\overline{\\rho}\\left(\\hat{\\mathbf{T}}_{\\phi}+r\\sin\\theta\\Omega\\mathbf{\\overline{U}^{m}}\\right)\\right)\\label{eq:angm}\\\\\n & + & \\boldsymbol{\\nabla\\cdot}\\left(r\\sin\\theta\\frac{\\overline{\\mathbf{B}}\\overline{B}_{\\phi}}{4\\pi}\\right).\\nonumber \n\\end{eqnarray}\nTo determine the meridional circulation we consider the azimuthal\ncomponent of the large-scale vorticity , $\\omega=\\left(\\boldsymbol{\\nabla}\\times\\overline{\\mathbf{U}}^{m}\\right)_{\\phi}$\n, which is governed by equation: \n\\begin{eqnarray}\n\\frac{\\partial\\omega}{\\partial t}\\!\\!\\! & \\negthinspace\\!=\\!\\!\\!\\! & r\\sin\\theta\\boldsymbol{\\nabla}\\cdot\\left(\\frac{\\hat{\\boldsymbol{\\phi}}\\times\\boldsymbol{\\nabla\\cdot}\\overline{\\rho}\\hat{\\mathbf{T}}}{r\\overline{\\rho}\\sin\\theta}-\\frac{\\mathbf{\\overline{U}}^{m}\\omega}{r\\sin\\theta}\\right)\\!\\!+r\\sin\\theta\\frac{\\partial\\Omega^{2}}{\\partial z}\\label{eq:vort}\\\\\n & +\\!\\!\\! & \\frac{1}{\\overline{\\rho}^{2}}\\left[\\boldsymbol{\\nabla}\\overline{\\rho}\\times\\boldsymbol{\\nabla}\\overline{p}\\right]_{\\phi}\\!\\!\\nonumber \\\\\n & + & \\!\\frac{1}{\\overline{\\rho}^{2}}\\left[\\!\\!\\boldsymbol{\\nabla}\\overline{\\rho}\\times\\left(\\!\\!\\boldsymbol{\\nabla}\\frac{\\overline{\\mathbf{B}}^{2}}{8\\pi}-\\frac{\\left(\\overline{\\mathbf{B}}\\boldsymbol{\\cdot\\nabla}\\right)\\overline{\\mathbf{B}}}{4\\pi}\\!\\right)\\!\\!\\right]_{\\phi},\\nonumber \n\\end{eqnarray}\nThe turbulent stresses tensor, $\\hat{\\mathbf{T}}$, is written in\nterms of small-scale fluctuations of velocity and magnetic field:\n\\begin{equation}\n\\hat{T}_{ij}=\\left(\\left\\langle u_{i}u_{j}\\right\\rangle -\\frac{1}{4\\pi\\overline{\\rho}}\\left(\\left\\langle b_{i}b_{j}\\right\\rangle -\\frac{1}{2}\\delta_{ij}\\left\\langle \\mathbf{b}^{2}\\right\\rangle \\right)\\right).\\label{eq:stres}\n\\end{equation}\nwhere ${\\partial\/\\partial z=\\cos\\theta\\partial\/\\partial r-\\sin\\theta\/r\\cdot\\partial\/\\partial\\theta}$\nis the gradient along the axis of rotation. The turbulent stresses\naffect generation and dissipation of large-scale flows, and they are\naffected by the global rotation and magnetic field. The magnitude\nof the kinetic coefficients in tensor $\\hat{\\mathbf{T}}$ depends\non the rms of the convective velocity, ${u}'$, the strength of the\nCoriolis force and the strength of the large-scale magnetic field.\nThe effect of the Coriolis force is determined by parameter $\\Omega^{*}=2\\Omega_{0}\\tau_{c}$,\nwhere $\\Omega_{0}=2.9\\times10^{-6}$rad\/s is the solar rotation rate\nand $\\tau_{c}$ is the convective turnover time. The effect of the\nlarge-scale magnetic field on the convective turbulence is determined\nby parameter $\\beta=\\left\\langle \\left|\\mathbf{B}\\right|\\right\\rangle \/\\sqrt{4\\pi\\overline{\\rho}u'^{2}}$.\n\nThe magnetic feedback on the coefficients of turbulent stress tensor\n$\\hat{\\mathbf{T}}$ was studied previously with the mean-field magnetohydrodynamic\nframework \\citep{rob-saw}. In our model we apply analytical results\nof \\citet{1994AN....315..157K}, \\citet{kit-rud-kuk} and \\citet{kuetal96}.\nThe analytical expression for $\\hat{\\mathbf{T}}$ is given in Appendix.\nIt was found that the standard components of the nondissipative of\n$\\hat{\\mathbf{T}}$ ($\\Lambda$-effect) are quenched with the increase\nof the magnetic field strength as $\\beta^{-2}$ and the magnetic quenching\nof the viscous parts is the order of $\\beta^{-1}$. Also, there is a non-trivial\neffect inducing the latitudinal angular momentum flux proportional\nto the magnetic energy \\citep{kit-rud-kuk,kuetal96}. This effect\nis quenched as $\\beta^{-2}$ for the case of the strong magnetic field.\nImplications of the magnetic feedback on the turbulent stress tensor\n$\\hat{\\mathbf{T}}$ were discussed in the models of solar torsional\noscillations and Grand activity cycles \\citep{kit-rud-kuk,kuetal96,p99,1999AA343.977K}.\nThe analytical results of the mean-field theory are in qualitative\nagreement with the direct numerical simulations \\citep{2007AN....328.1006K,kap2011,2017arXiv171208045K}.\n\nProfile of $\\tau_{c}$ (as well as profiles of $\\overline{\\rho}$\nand other thermodynamic parameters) is obtained from a standard solar\ninterior model calculated using the MESA code \\citep{mesa11,mesa13}.\nThe rms velocity, $u'$, is determined in the mixing length approximations\nfrom the gradient of the mean entropy, $\\overline{s}$, \n\\begin{equation}\nu'=\\frac{\\ell}{2}\\sqrt{-\\frac{g}{2c_{p}}\\frac{\\partial\\overline{s}}{\\partial r}},\\label{eq:uc}\n\\end{equation}\nwhere $\\ell=\\alpha_{MLT}H_{p}$ is the mixing length, $\\alpha_{MLT}=2.2$\nis the mixing length theory parameter, and $H_{p}$ is the pressure\nscale height. For a non-rotating star the ${u}'$ profile corresponds\nto results of the MESA code. The mean-field equation for heat transport\ntakes into account effects of rotation and magnetic field \\citep{2000ARep...44..771P}:\n\\begin{equation}\n\\overline{\\rho}\\overline{T}\\left(\\frac{\\partial\\overline{s}}{\\partial t}+\\left(\\overline{\\mathbf{U}}\\cdot\\boldsymbol{\\nabla}\\right)\\overline{s}\\right)=-\\boldsymbol{\\nabla}\\cdot\\left(\\mathbf{F}^{conv}+\\mathbf{F}^{rad}\\right)-\\hat{T}_{ij}\\frac{\\partial\\overline{U}_{i}}{\\partial r_{j}}-\\frac{1}{4\\pi}\\boldsymbol{\\mathcal{E}}\\cdot\\nabla\\times\\boldsymbol{\\overline{B}},\\label{eq:heat}\n\\end{equation}\nwhere, $\\overline{\\rho}$ and $\\overline{T}$ are the mean density\nand temperature, $\\boldsymbol{\\mathcal{E}}=\\left\\langle \\mathbf{u\\times b}\\right\\rangle $\nis the mean electromotive force. The Eq.(\\ref{eq:heat}) includes\nthe thermal energy loss and gain due to generation and dissipation\nof large-scale flows. The last term of the Eq.(\\ref{eq:heat}) takes\ninto account effect of thermal energy exchange because of dissipation\nand generation of magnetic field \\citep{2000ARep...44..771P}. In\nderivation of the mean-field heat transport equation (see, \\citealp{2000ARep...44..771P}),\nit was assumed that the magnetic and rotational perturbations of the\nreference thermodynamic state are small. Also the parameters of the\nreference state are given independently by the MESA code.\n\nFor the anisotropic convective flux we employ the expression suggested\nby \\citet{1994AN....315..157K} (hereafter KPR94), \n\\begin{equation}\nF_{i}^{conv}=-\\overline{\\rho}\\overline{T}\\chi_{ij}\\nabla_{j}\\overline{s}.\\label{conv}\n\\end{equation}\nThe further details about dependence of the eddy conductivity tensor\n$\\chi_{ij}$ from effects of both the global rotation and large-scale\nmagnetic field are given in Appendix. The diffusive heat transport\nby radiation reads, \n\\[\n\\mathbf{F}^{rad}=-c_{p}\\overline{\\rho}\\chi_{D}\\boldsymbol{\\nabla}T,\n\\]\nwhere \n\\[\n\\chi_{D}=\\frac{16\\sigma\\overline{T}^{3}}{3\\kappa\\overline{\\rho}^{2}c_{p}}.\n\\]\nBoth the eddy conductivity and viscosity are determined from the mixing-length\napproximation: \n\\begin{eqnarray}\n\\chi_{T} & = & \\frac{\\ell^{2}}{4}\\sqrt{-\\frac{g}{2c_{p}}\\frac{\\partial\\overline{s}}{\\partial r}},\\label{eq:chi}\\\\\n\\nu_{T} & = & \\mathrm{Pr_{T}}\\chi_{T},\\label{eq:nu}\n\\end{eqnarray}\nwhere $\\mathrm{Pr_{T}}$ is the turbulent Prandtl number. {Note,\nthat in Eq\\eqref{eq:chi} we employ factor $1\/2$ instead of $1\/3$.\nWith this choice the distribution of the mean entropy gradient, which\nresults from solution of the Eq\\eqref{eq:heat} for the nonrotating\nand nonmagnetic case is close to results of the MESA code.} It is\nassumed that $\\mathrm{Pr_{T}}={\\displaystyle 3\/4}$. {This\ncorresponds to the theoretical results of KPR94}. For this choice\nwe have the good agreement with solar angular velocity latitudinal\nprofile. We assume that the solar rotation rate corresponds to rotation\nrate of solar tachocline at 30$^{\\circ}$ latitude, i.e., $\\Omega_{0}\/2\\pi=430$nHz\n\\citep{1997SoPh170.43K}. We employ the stress-free boundary conditions\nin the hydrodynamic part of the problem. For the Eq(\\ref{eq:heat})\nthe thermal flux at the bottom is taken from the MESA code. At the\ntop, the thermal flux from the surface is approximated by the flux\nfrom a blackbody: \n\\begin{equation}\nF_{r}=\\frac{L_{\\odot}}{4\\pi r^{2}}\\left(1+4\\frac{T_{e}}{T_{eff}}\\frac{\\overline{s}}{c_{p}}\\right),\\label{eq:flx}\n\\end{equation}\nwhere where $T_{eff}$ is the effective temperature of the photosphere\nand $T_{e}$ is the temperature at the outer boundary of the integration\ndomain.\n\nFigure \\ref{fig:sun-flow} shows profiles of the angular velocity,\nstreamlines of the meridional circulations and the radial profiles\nof the angular velocity and the meridional flow velocity for a set\nof latitudes. The given results were discussed in details by \\citet{2018ApJ854.67P}.\nThe model shows the double-layer circulation pattern with the upper\nstagnation point at $r=0.88R_{\\odot}$. The amplitude of the surface\npoleward flow is about 15m\/s. The angular velocity profile shows a\nstrong subsurface shear that is higher at low latitudes and it is\nless near poles. Contrary to results of \\citet{Zhao13m} and model\nof \\citet{PK13} the double-cell meridional circulation structure\nextends from equator to pole. This is partly confirmed by the new\nresults of helioseismology by \\citet{2017ApJ849.144C} who also found\nthat the poleward flow at the surface goes close enough to pole. {It\nis important no mention that the current results of the helioseismic\ninversions for the meridional circulation remains controversial For\nexample, \\citet{2015ApJ813.114R} found that the meridional circulation\ncan be approximated by a single-cell structure with the return flow\ndeeper than 0.77R$_{\\odot}$. However, their results indicate an additional\nweak cell in the equatorial region, and contradict to the recent results\nof \\citet{2017ApJ845.2B} who confirmed a shallow return flow at 0.9R$_{\\odot}$.\nAlso, their results indicated that the upper meridional circulation\ncell extends close to the solar pole.}\n\n\\begin{figure}\n\\includegraphics[width=0.8\\columnwidth]{fig1}\n\n\\caption{\\label{fig:sun-flow}a) angular velocity profile, $\\Omega\\left(r,\\theta\\right)\/2\\pi$,\ncontours are in range of 327-454 nHz; b) the radial profiles of the\nangular velocity for latitudes: $\\varphi=0^{\\circ}$, $30^{\\circ}$\nand $60^{\\circ}$; c) streamlines of the meridional circulation; d)\nradial profile of the meridional flow at $\\theta=45^{\\circ}$.}\n\\end{figure}\n\n\n\\subsection{Dynamo equations}\n\nWe model evolution of the large-scale axisymmetric magnetic field,\n$\\overline{\\mathbf{B}}$, by the mean-field induction equation \\citep{KR80},\n\\begin{equation}\n\\partial_{t}\\overline{\\mathbf{B}}=\\boldsymbol{\\nabla}\\times\\left(\\boldsymbol{\\mathcal{E}}+\\mathbf{\\overline{U}}\\times\\overline{\\mathbf{B}}\\right),\\label{eq:mfe-1}\n\\end{equation}\nwhere, $\\boldsymbol{\\mathcal{E}}=\\left\\langle \\mathbf{u\\times b}\\right\\rangle $\nis the mean electromotive force with $\\mathbf{u}$ and $\\mathbf{b}$\nstanding for the turbulent fluctuating velocity and magnetic field\nrespectively.\n\nSimilar to our recent paper (see, \\citep{2014ApJ_pipk,2017MNRAS.466.3007P}),\nwe employ the mean electromotive force in form: \n\\begin{equation}\n\\mathcal{E}_{i}=\\left(\\alpha_{ij}+\\gamma_{ij}\\right)\\overline{B}_{j}-\\eta_{ijk}\\nabla_{j}\\overline{B}_{k}.\\label{eq:EMF-1-1}\n\\end{equation}\nwhere symmetric tensor $\\alpha_{ij}$ models the generation of magnetic\nfield by the $\\alpha$- effect; antisymmetric tensor$\\gamma_{ij}$\ncontrols the mean drift of the large-scale magnetic fields in turbulent\nmedium, including the magnetic buoyancy; the tensor $\\eta_{ijk}$\ngoverns the turbulent diffusion. The reader can find further details\nabout the $\\boldsymbol{\\mathcal{E}}$ in the above cited papers.\n\nThe $\\alpha$ effect takes into account the kinetic and magnetic helicities\nin the following form: \n\\begin{eqnarray}\n\\alpha_{ij} & = & C_{\\alpha}\\eta_{T}\\psi_{\\alpha}(\\beta)\\alpha_{ij}^{(H)}+\\alpha_{ij}^{(M)}\\frac{\\overline{\\chi}\\tau_{c}}{4\\pi\\overline{\\rho}\\ell^{2}},\\label{alp2d-2}\\\\\n\\eta_{T} & = & \\frac{\\nu_{T}}{\\mathrm{Pm_{T}}}\n\\end{eqnarray}\nwhere $C_{\\alpha}$ is a free parameter which controls the strength\nof the $\\alpha$- effect due to turbulent kinetic helicity; tensors\n$\\alpha_{ij}^{(H)}$ and $\\alpha_{ij}^{(M)}$ express the kinetic\nand magnetic helicity parts of the $\\alpha$-effect, respectively;\n$\\mathrm{Pm_{T}}$ is the turbulent magnetic Prandtl number, and $\\overline{\\chi}=\\left\\langle \\mathbf{a}\\cdot\\mathbf{b}\\right\\rangle $\n($\\mathbf{a}$ and $\\mathbf{b}$ are the fluctuating parts of magnetic\nfield vector-potential and magnetic field vector). Both the $\\alpha_{ij}^{(H)}$\nand the $\\alpha_{ij}^{(M)}$ depend on the Coriolis number. Function\n$\\psi_{\\alpha}(\\beta)$ controls the so-called ``algebraic'' quenching\nof the $\\alpha$- effect where $\\beta=\\left|\\overline{\\mathbf{B}}\\right|\/\\sqrt{4\\pi\\overline{\\rho}u'^{2}}$,\n$u'$ is the RMS of the convective velocity. It is found that $\\psi_{\\alpha}(\\beta)\\sim\\beta^{-3}$\nfor $\\beta\\gg1$. The $\\alpha$- effect tensors $\\alpha_{ij}^{(H)}$\nand $\\alpha_{ij}^{(M)}$ are given in Appendix. \n\\begin{figure}\n\\includegraphics[width=0.9\\columnwidth]{fig2}\n\n\\caption{\\label{fig:alp} a) the radial profiles of the total (solid line)\nand anisotropic (dashed line) parts of the eddy diffusivity at $\\theta=45^{\\circ}$;\nb) radial profiles of the kinetic $\\alpha$-effect components at $\\theta=45^{\\circ}$;\nc) the equipartition strength of the magnetic field, $B_{eq}=\\sqrt{4\\pi\\overline{\\rho}u'^{2}}$,\nwhere $u'$ is determined by the equatorial profile of the mean entropy,\nsee the Eq(\\ref{eq:uc}); the dashed line is from results of the reference\nmodel (MESA code) and the solid line is for the rotating convection\nzone,i.e., after solution of the Eq(\\ref{eq:heat}).}\n\\end{figure}\n\nContribution of the magnetic helicity to the $\\alpha$-effect is expressed\nby the second term in Eq.(\\ref{alp2d-2}). The evolution of the turbulent\nmagnetic helicity density, $\\overline{\\chi}=\\left\\langle \\mathbf{a}\\cdot\\mathbf{b}\\right\\rangle $,\nis governed by the conservation law \\citep{pip13M}:\n\n\\begin{eqnarray}\n\\frac{\\partial\\overline{\\chi}}{\\partial t} & = & -2\\left(\\boldsymbol{\\mathcal{E}}\\cdot\\overline{\\bm{B}}\\right)-\\frac{\\overline{\\chi}}{R_{m}\\tau_{c}}+\\boldsymbol{\\nabla}\\cdot\\left(\\eta_{\\chi}\\boldsymbol{\\nabla}\\bar{\\chi}\\right)\\label{eq:hel-1}\\\\\n & & -\\eta\\overline{\\mathbf{B}}\\cdot\\mathbf{\\overline{J}}-\\boldsymbol{\\nabla}\\cdot\\left(\\boldsymbol{\\mathcal{E}}\\times\\overline{\\mathbf{A}}\\right),\\nonumber \n\\end{eqnarray}\nwhere $R_{m}=10^{6}$ is the magnetic Reynolds number and $\\eta$\nis the microscopic magnetic diffusion. In the drastic difference to\nanzatz of \\citet{kleruz82}, the Eq(\\ref{eq:hel-1}) contains the\nterm $\\left(\\boldsymbol{\\mathcal{E}}\\times\\overline{\\mathbf{A}}\\right)$.\nIt consists of the magnetic helicity density fluxes which result from\nthe large-scale magnetic dynamo wave evolution. The given contribution\nalleviates the catastrophic quenching problem \\citep{hub-br12,pip13M}.\nAlso the catastrophic quenching of the $\\alpha$-effect can be alleviated\nwith help of the diffusive flux of the turbulent magnetic helicity,\n$\\boldsymbol{\\boldsymbol{\\mathcal{F}}}^{\\chi}=-\\eta_{\\chi}\\boldsymbol{\\nabla}\\bar{\\chi}$\n\\citep{guero10,chatt11}. The coefficient of the turbulent helicity\ndiffusivity, $\\eta_{\\chi}$, is a parameter in our study. It affects\nthe hemispheric helicity transfer \\citep{mitra10}.\n\nIn the model we take into account the mean drift of large-scale field\ndue to the magnetic buoyancy, $\\gamma_{ij}^{(buo)}$ and the gradient\nof the mean density, $\\gamma_{ij}^{(\\Lambda\\rho)}$: \n\\begin{eqnarray}\n\\gamma_{ij} & = & \\gamma_{ij}^{(\\Lambda\\rho)}+\\gamma_{ij}^{(buo)},\\nonumber \\\\\n\\gamma_{ij}^{(\\Lambda\\rho)} & = & 3C_{pum}\\eta_{T}\\left(f_{1}^{(a)}\\left(\\mathbf{\\boldsymbol{\\Omega}}\\cdot\\boldsymbol{\\Lambda}^{(\\rho)}\\right)\\frac{\\Omega_{n}}{\\Omega^{2}}\\varepsilon_{inj}-\\frac{\\Omega_{j}}{\\Omega^{2}}\\varepsilon_{inm}\\Omega_{n}\\Lambda_{m}^{(\\rho)}\\right)\\label{eq:pump1}\\\\\n\\gamma_{ij}^{(buo)} & = & -\\frac{\\alpha_{MLT}u'}{\\gamma}\\beta^{2}K\\left(\\beta\\right)g_{n}\\varepsilon_{inj},\\nonumber \n\\end{eqnarray}\nwhere $\\mathbf{\\boldsymbol{\\Lambda}}^{(\\rho)}=\\boldsymbol{\\nabla}\\log\\overline{\\rho}$\n; functions $f_{1}^{(a)}$ and $K\\left(\\beta\\right)$ are given in\n\\cite{kp93,2017MNRAS.466.3007P}. The standard choice of the pumping\nparameter is $C_{pum}=1$. In this case the pumping velocity is scaled\nin the same way as the magnetic eddy diffusivity. In the presence\nof the multi-cell meridional circulation, the direction and magnitude\nof the turbulent pumping become critically important for the modelled\nevolution of the magnetic field. It is confirmed in the direct numerical\nsimulations, as well (see, \\cite{2018AA609A..51W}). For the standard\nchoice, the turbulent pumping is about an order of magnitude less than\nthe meridional circulation. For this case, explanation of the latitudinal\ndrift of the toroidal magnetic field near the surface faces a problem\n(cf., \\cite{PK13}). To study the effect of turbulent pumping we\nintroduce this parameter $C_{pum}$.\n\nFor the bottom boundary we apply the perfect conductor boundary conditions:\n$\\mathcal{E}_{\\theta}=0,\\,A=0$. The boundary conditions at the top\nare defined as follows. Firstly, following ideas of \\citet{1992AA256371M}\nand \\citet{pk11apjl} we formulate the boundary condition in the form\nthat allows penetration of the toroidal magnetic field to the surface:\n\\begin{eqnarray}\n\\delta\\frac{\\eta_{T}}{r_{e}}B+\\left(1-\\delta\\right)\\mathcal{E}_{\\theta} & = & 0,\\label{eq:tor-vac}\n\\end{eqnarray}\nwhere $r_{e}=0.99R_{\\odot}$, and parameter $\\delta=0.99$. The magnetic\nfield potential in the outside domain is \n\\begin{equation}\nA^{(vac)}\\left(r,\\mu\\right)=\\sum a_{n}\\left(\\frac{r_{e}}{r}\\right)^{n}\\sqrt{1-\\mu^{2}}P_{n}^{1}\\left(\\mu\\right).\\label{eq:vac-dec}\n\\end{equation}\n\nThe coupled angular momentum and dynamo equations are solved using\nfinite differences for integration along the radius and the pseudospectral\nnodes for integration in latitude. The number of mesh points in radial\ndirection was varied from $100$ to $150$. The nodes in latitude\nare zeros of the Legendre polynomial of degree ${N}$, where N was\nvaried from ${N=64}$ to ${N=84}$. The resolution with 64 nodes in\nlatitude and with 100 points in radius was found satisfactory. The\nmodel employed the Crank-Nicolson scheme, using a half of the time-step\nfor integration in the radial direction and another half for integration\nalong latitude.\n\nTo quantify the mirror symmetry type of the toroidal magnetic field\ndistribution relative to equator we introduce the parity index $P$:\n\\begin{eqnarray}\nP & = & \\frac{E_{q}-E_{d}}{E_{q}+E_{d}},\\label{eq:parity}\\\\\nE_{d} & = & \\int\\left(B\\left(r_{0},\\theta\\right)-B\\left(r_{0},\\pi-\\theta\\right)\\right)^{2}\\sin\\theta d\\theta,\\nonumber \\\\\nE_{q} & = & \\int\\left(B\\left(r_{0},\\theta\\right)+B\\left(r_{0},\\pi-\\theta\\right)\\right)^{2}\\sin\\theta d\\theta,\\nonumber \n\\end{eqnarray}\nwhere $E_{d}$ and $E_{q}$ are the energies of the dipole-like and\nquadruple-like modes of the toroidal magnetic field at $r_{0}=0.9R_{\\odot}$.\nAnother integral parameter is the mean density of the toroidal magnetic\nfield in the subsurface shear layer: \n\\begin{equation}\n\\overline{B^{T}}=\\sqrt{E_{d}+E_{q}}.\\label{eq:bt}\n\\end{equation}\nAnother parameter characterize the mean strength of the dynamo processes\nin the convection zone: \n\\begin{equation}\n\\overline{\\beta}=\\left\\langle \\left|\\overline{\\mathbf{B}}\\right|\/\\sqrt{4\\pi\\overline{\\rho}u'^{2}}\\right\\rangle ,\\label{eq:bet}\n\\end{equation}\nwhere the averaging is done over the convection zone volume. The boundary\nconditions Eq(\\ref{eq:tor-vac}) provide the Poynting flux of the\nmagnetic energy out of the convection zone. Taking into account the\nEq(\\ref{eq:flx}) the variation of the thermal flux at the surface\nare given as follows: \n\\begin{eqnarray}\n\\delta F & = & \\delta F_{c}+\\delta F_{B}\\label{eq:dF}\\\\\n\\delta F_{c} & = & 4\\frac{T_{e}}{T_{eff}}\\frac{\\delta\\overline{s}}{c_{p}}\\label{eq:flxb}\\\\\n\\delta F_{B} & = & \\frac{1}{4\\pi}\\left(\\mathcal{E}_{\\phi}\\overline{B}_{r}-\\mathcal{E}_{\\theta}\\overline{B}_{\\phi}\\right),\\label{eq:fcfb}\n\\end{eqnarray}\nwhere $\\delta\\overline{s}$ is the entropy variation because of the\nmagnetic activity. The second term of the Eq(\\ref{eq:dF}) governs\nthe magnetic energy input in the stellar corona.\n\n\\section{Results}\n\nTo match the solar cycle period we put $\\mathrm{Pm_{T}}=10$ in all\nour models. {The theoretical estimations of \\citet{1994AN....315..157K}\ngives $\\mathrm{Pm_{T}}=4\/3$. This is the long standing theoretical\nproblem of the solar dynamo period \\citep{brsu05}. Currently, the\nsolar dynamo period can be reproduced for $\\mathrm{Pm_{T}}\\gg1$.\nThe issue exists both in the distributed and in the flux-transport\ndynamo. Moreover, the flux-transport dynamo can reproduce the observation\nonly with the special radial profile of the eddy diffusivity (see,\ne.g., \\citep{2006ApJ...647..662R}). In our models we employ the rotational\nquenching the eddy diffusivity coefficients and the high $\\mathrm{Pm_{T}}$.\nFigures \\ref{fig:alp}a and b show the radial profiles of the eddy\ndiffusivity coefficients and components of the $\\alpha_{ij}^{(H)}$\nat latitude $45^{\\circ}$ in our model for $\\mathrm{Pm_{T}}=10$.\nIn the upper part of the convection zone the magnitude of the turbulent\nmagnetic diffusivity is close to estimations of \\citet{1993AA...274..521M}\nbased on observations of the sunspot decay rate. The eddy diffusivity\nis an order of $10^{10}$cm\/s and less near the bottom of the convection\nzone. The diffusivity profile is the same as in our previous paper\n\\citet{2014ApJ_pipk}. }\n\n{The radial profiles of the $\\alpha$ effect for $C_{\\alpha}=C_{\\alpha}^{(cr)}$\nare illustrated in Figure \\ref{fig:alp}b. The $\\alpha$ effect (cf,\nthe above discussion about $\\Lambda$-effect) change the sign near\nthe bottom of the convection zone. This is also found in the direct\nnumerical simulations \\citep{2006AA455.401K}.}\n\nTable \\ref{tab:C} gives the list of our models, their control and\noutput parameters. We sort the models with respect to magnitude of\nthe $\\alpha$ effect using the ratio ${\\displaystyle \\frac{C_{\\alpha}}{C_{\\alpha}^{(cr)}}}$,\nthe magnitude of the eddy-diffusivity of the magnetic helicity density,\n${\\displaystyle \\frac{\\eta_{\\chi}}{\\eta_{T}}}$, where $\\eta_{T}=\\nu_{T}\/\\mathrm{Pm_{T}}$,\nand $\\nu_{T}$ is determined from Eq(\\ref{eq:nu}), and with respect\nof the magnetic feedback on the differential rotation.{ The\n$C_{pum}$ controls the pumping velocity magnitude (see, Eq\\eqref{eq:pump1});\nthe parameter ${\\displaystyle \\frac{\\Delta\\Omega}{\\Omega_{0}}}$ show\nthe relative difference of the surface angular velocity between the\nsolar equator and pole; the strength of the dynamo is characterized\nby the range of the magnetic cycle variations of $\\overline{\\beta}$\n(see, Eq\\eqref{eq:bet}); the dynamo cycle period; the magnitude of\nthe surface meridional circulation. From the Table 1 we see that the\nnonkinematic runs show the magnetic cycle variations of ${\\displaystyle \\frac{\\Delta\\Omega}{\\Omega_{0}}}$\nand the surface meridional circulation. }\n\nFigure \\ref{fig:alp}b shows the radial profiles of the equipartition\nstrength of the magnetic field, $B_{eq}=\\sqrt{4\\pi\\overline{\\rho}u'^{2}}$\nin the solar convection zone for the reference model (non-rotating)\ngiven by MESA code and in the rotating convective zone. In the rotating\nconvection zone, the mean-entropy gradient is larger than in the nonrotating\ncase. {This is because of the rotational quenching of the eddy-conductivity.\nThe magnitude of the convective heat flux is determined by the boundary\ncondition at the bottom of the convection zone and it remains the\nsame for the rotating (our model) and nonrotating (MESA code) cases.\nAssuming that the convective turnover time is not subjected to the\nrotational quenching, the reduction of the eddy conductivity because\nof the rotational quenching is compensated by the increase of the\nmean-entropy gradient.} This results in the increase of the parameter\n$B_{eq}$.\n\n{The increase of the RMS convective velocity in case of the\nrotating convection zone seems to contradict the results of direct numerical\nsimulations of \\cite{2016AA596A.115W}. This is\nlikely because of inconsistent assumptions behind the MLT expression\nfor the RMS convective velocity, see Eq(\\ref{eq:uc}). The given issue\ncan affect the amplitude of the dynamo generated magnetic field near the\nbottom of the convection zone. Our models operate in regimes where\n$\\left|B\\right|\\le B_{eq}$, and the substantial part of the dynamo\nquenching is due to magnetic helicity conservation. Therefore the\ngiven issue does not much affect our results.}\n\n\n\n\\begin{table}\n\\caption{\\label{tab:C}Control and output parameters of the dynamo models.}\n\\begin{tabular}{>{\\centering}p{1cm}>{\\centering}p{2cm}>{\\centering}p{1cm}>{\\centering}p{2cm}>{\\centering}p{1.5cm}>{\\centering}p{2cm}>{\\centering}p{1cm}>{\\centering}p{1cm}}\n\\toprule \nModel & $C_{pum}$ & ${\\displaystyle \\frac{\\eta_{\\chi}}{\\eta_{T}}}$ & ${\\displaystyle \\frac{C_{\\alpha}}{C_{\\alpha}^{(cr)}}}$ & ${\\displaystyle \\frac{\\Delta\\Omega}{\\Omega_{0}}}$ & $\\overline{\\beta}$ & \\begin{centering}\nPeriod \n\\par\\end{centering}\n{[}YR{]} & $\\max U_{\\theta}$\n\n{[}M\/S{]}\\tabularnewline\n\\midrule \nM1 & 1 & $0.1$ & 1.1 & 0.279 & 0.05-0.1 & 6 & 15\\tabularnewline\n\\midrule \nM2 & Pm$_{T}$ & 0.1 & 1.1 & 0.279 & 0.13-0.24 & 10.5 & 15\\tabularnewline\n\\midrule \nM4 & -\/- & 0.3 & -\/- & 0.279 & 0.15-0.3 & 10.5,12.05 & 15\\tabularnewline\n\\midrule \nM5 & -\/- & 0.01 & -\/- & 0.279 & 0.14-0.26 & 9.3 & 15\\tabularnewline\n\\midrule \n & & & Nonkinematic & runs & & & \\tabularnewline\n\\midrule \nM3 & Pm$_{T}$ & 0.1 & 1.1 & 0.263-0.275 & 0.11-0.21 & 10.3 & 15.0\n\n$\\pm0.5$\\tabularnewline\n\\midrule \nM3a2 & -\/- & -\/- & 2 & 0.232-0.253 & 0.36-0.66 & 4.7,265 & 14.0\n\n$\\pm2.1$\\tabularnewline\n\\midrule \nM3a3 & -\/- & -\/- & 3 & 0.22-0.24 & 0.52-0.91 & 4.0 & 14.5 $\\pm2.5$\\tabularnewline\n\\midrule \nM3a4 & -\/- & -\/- & 4 & 0.21-0.245 & 0.68-1.05 & 3.4 & 14.7\n\n$\\pm3.$5\\tabularnewline\n\\bottomrule\n\\end{tabular}\n\\end{table}\n\n\n\\subsection{Effects of turbulent pumping}\n\nAs the first step, we consider the kinematic dynamo model with the\nnonlinear $\\alpha$ effect. Results of the model M1 are shown in Figure\n\\ref{pumpa}. The model M1 roughly agree with results of \\citet{PK13}\n(hereafter, PK13). It employs the same mean electromotive force as\nin our previous paper. In particular, the maximum pumping velocity\nis the order of 1m\/s. The effective velocity drift due to the magnetic\npumping and meridional circulation is shown in Figures \\ref{pumpa}(a)\nand (b). Figures \\ref{pumpa}(d) and (e) show the time-latitude variations\nof the toroidal magnetic field at $r=0.9R$ and in the middle of the\nconvection zone. The agreement with the solar observations is worse\nthan in the previous model PK13 because of difference in the meridional\ncirculation structure. The model of PK13 employed the meridional circulation\nprofile provided by results of \\citet{Zhao13m}. In that profile,\nthe near-surface meridional circulation cell is more shallow and it\ndoes not touch the pole as it happens in the present model, see, Figure\n\\ref{fig:sun-flow}. By this reason, the polar magnetic field in model\nM1 is much larger than in results of PK13.\n\n\\begin{figure}\n\\includegraphics[width=0.8\\columnwidth]{m1}\n\n\\caption{\\label{pumpa}a) Direction of pumping velocity of the toroidal magnetic\nfield in model M1; b) the effective velocity drift of the toroidal\nmagnetic field (pumping + meridional circulation); c) the snapshot\nof the toroidal magnetic field distribution (color image) and streamlines\nof the poloidal magnetic field in the Northern hemisphere of the Sun;\nd) the time-latitude diagram of the toroidal magnetic field evolution\n(contours in range of of $\\pm500$G at $r=0.9R$ and radial magnetic\nfield at the surface (color image).}\n\\end{figure}\n\nFor the purpose of our study, it is important to get the properties\nof the dynamo solution as close as possible to results of solar observations.\nTo solve the above issues we increase the turbulent pumping velocity\nmagnitude by factor $\\mathrm{Pm_{T}}$. The results are shown in Figure\n\\ref{pumpb}. The model has the correct time-latitude diagram of the\ntoroidal magnetic field in the subsurface shear layer. The surface\nradial magnetic field evolves in agreement with results of observations\n\\citep{2013AARv2166S}. The magnitude of the polar magnetic field\nis 10 G, which is in a better agreement with observations (e.g., \\cite{2007AAS...210.2405L})\nthan the model M1. Figures \\ref{pumpb} (a) and (b) show the effective\nvelocity drift of the large-scale toroidal magnetic field. The equatorward\ndrift with magnitude the order of 1-2 m\/s operates in major part of the\nsolar convection zone from $0.75R$ to $0.91R$. Interesting that\nthe obtained results are similar to those from the direct numerical\nsimulation of \\citet{2018AA609A..51W}. Note that in the given model\nthe magnitude of the pumping velocity is about factor 2 less than\nin results of \\citet{2018AA609A..51W}. It seems that some of the\nissues in model M1 would be less pronounced if the meridional circulation\npattern was closer to results of helioseismology of \\citet{Zhao13m}\nor \\citet{2017ApJ849.144C}. However results of the direct numerical\nsimulations of \\citep{2016AA596A.115W,2018AA609A..51W} seem to show\nthat the evolution of the large-scale magnetic field inside convection\nzone does not depend much on the meridional circulation. This argues\nfor the strong magnetic pumping effects in the global dynamo. This\nimportant issue can be debated further. This is out of the main scopes\nof this paper. The rest of our models employ the same pumping effect\nas in the model M2 (see, Table \\ref{tab:C}). \n\n\\begin{figure}\n\\includegraphics[width=0.8\\columnwidth]{m2}\n\n\\caption{\\label{pumpb}a) Direction of pumping velocity of the toroidal magnetic\nfield in model M2; b) the effective velocity drift of the toroidal\nmagnetic field (pumping + meridional circulation); c) the snapshot\nof the toroidal magnetic field distribution (color image) and streamlines\nof the poloidal magnetic field in the Northern hemisphere of the Sun;\nd) the time-latitude diagram of the toroidal magnetic field evolution\n(contours in range of of $\\pm1$kG at $r=0.9R$ and radial magnetic\nfield at the surface (color image).}\n\\end{figure}\n\n\n\\subsection{The global flows oscillations in magnetic cycle}\n\n\\begin{figure}\n\\includegraphics[width=0.8\\columnwidth]{m3}\n\n\\caption{\\label{fig:s-bat}The model M3, a) Time-latitude butterfly diagram\nfor the toroidal field in the upper part of the convection zone (color\nimage) and the surface radial magnetic field shown by contours ($\\pm$5G);\nb) the surface variations of the azimuthal velocity (color image)\nand the meridional velocity (contours in the range of $\\pm0.5$ m\/s). }\n\\end{figure}\n\nFigure \\ref{fig:s-bat} show the time-latitude diagrams of the magnetic\nfield and the global flow variations for the model M3. The torsional\noscillations on the surface are about $\\pm2$m\/s. They are defined\nas follows, $\\delta U_{\\phi}=\\left(\\Omega\\left(r,\\theta t\\right)-\\overline{\\Omega\\left(r,\\theta,t\\right)}\\right)r\\sin\\theta$,\nwhere the averaging is done over the stationary phase of evolution.\nThe torsional wave has both the equator- and poleward branches. In\nthe equatorward torsional wave, the change from the positive to negative\nvariation goes about 2 years ahead of the maxima of the toroidal magnetic\nfield wave. This agrees with results of observations of \\citet{2011JPhCS271a2074H}\nand with direct numerical simulations of \\citet{2016ApJ828L.3G}.\nThe magnitude of the meridional flow variations agrees with results\nof \\citet{2014ApJ789L7Z}. Also, we see that on the surface the meridional\nvelocity variations converge toward the maximum of the toroidal magnetic\nfield wave. This is also in qualitative agreement with the observations.\n\n\\begin{figure}\n\\includegraphics[width=0.95\\columnwidth]{m3s}\\caption{\\label{fig:M3}The model M3: a) snapshots of the magnetic field in\nfour phase of the magnetic cycle, the toroidal magnetic field strength\nis shown by color, contours show streamlines of the poloidal magnetic\nfield; b) color image show variations of the angular velocity, contours\n(range of $\\pm0.5$m\/s) show variations of the meridional flow; c)\ncontours show the azimuthal component of the total (kinetic and magnetic\nhelicity parts)$\\alpha$ -effect, the background image shows the part\nof the $\\alpha_{\\phi\\phi}$ induced by the magnetic helicity conservation\n(see, the second term of Eq(\\ref{alp2d-2})).}\n\\end{figure}\n\nFigure \\ref{fig:M3} shows snapshots of the magnetic field, the global\nflows variations and the azimuthal component of the total (kinetic\nand magnetic helicity parts)$\\alpha$ -effect for a half magnetic\ncycle. The Figure shows that a new cycle starts at the bottom of the\nconvection zone. The main part of the dynamo wave drifts to surface\nequatorward. There is a polar branch which propagates poleward along\nthe bottom of the convection zone. The torsional oscillations, as\nwell as, the meridional flow variations are elongated along the axis\nof rotation. This can be interpreted as a result of mechanical perturbation\nof the Taylor-Proudman balance \\citep{2006ApJ...647..662R}. We postpone\nthe detailed analysis of the torsional oscillation to another paper.\n{Figure \\ref{fig:M3}b shows that maxima of the meridional\nflow variations are located at the upper boundary of the dynamo domain.}\nThis is because the main drivers of the meridional circulation, which\nare the baroclinic forces, have the maximum near the boundaries of\nthe solar convection zone \\citep{rem2005ApJ,2015ApJ...804...67F,2017arXiv170202421P}.\nIn comparing Figures \\ref{fig:M3}c and \\ref{fig:alp} it is seen\nthat the dynamo wave affect the $\\alpha$ -effect. Also, in agreement\nwith our previous model \\citep{pip2013ApJ}, we find that the magnetic\nhelicity conservation results into increasing the $\\alpha$ -effect\nin the subsurface shear layer. It occurs just ahead of the dynamo\nwave drifting toward the top. The given effect support the equatorward\npropagation of the large-scale toroidal field in subsurface shear\nlayer \\citep{kap2012}.\n\nThe increasing the $\\alpha$-effect parameter results in a number\nof consequences for the non-linear evolution of the large-scale magnetic\nfield. The dynamo period is decreasing with the increase of the $\\alpha$-effect\n\\citep{pk11}. The magnitude of the dynamo wave increases with the\nincrease of the parameter $C_{\\alpha}$.{ Therefore, our models\nshow that in the distributed solar-type dynamo the dynamo period can\ndecrease with the increase of the magnetic activity level. This is\nin agreement with the results of the stellar activity observations\nof \\citet{1984ApJ...287..769N,2009AA_strassm,2017PhDT3E}. Here we\nfor the first time demonstrate this effect in the distributed dynamo\nmodel with the meridional circulation. }Figure \\ref{fig:m8} shows\nresults for the model M3a4 with ${\\displaystyle C_{\\alpha}=4C_{\\alpha}^{(cr)}}$.\nThe model shows the solar-like dynamo waves in the subsurface shear\nlayer. The toroidal magnetic field reaches the strength of 3kG in\nthe upper part of the convection zone. Simultaneously, the polar magnetic\nfield has the maximum strength of 100 G. Variations of the zonal and\nmeridional flows on the surface are about of factor 6 larger than\nin the model M3. The model M3a4 show a high level magnetic activity\nwith a strong toroidal magnetic field in the subsurface layer and\nvery strong polar field. {Results of stellar observations show\nthat this is expected on the young solar analogs and late K-dwarfs\nas well (e.g., \\citep{2005LRSP2.8B,2016MNRAS1129S}). However, the\ngiven results can not be consistent with those cases because in our\nmodel the rotation rate is much slower than for the young solar-type\nstars. Results of the linear models show that the internal differential\nrotation and meridional circulation change with increase of the stellar\nangular velocity \\citep{kit11}.}\n\n\\begin{figure}\n\\includegraphics[width=0.8\\columnwidth]{m8}\n\n\\caption{\\label{fig:m8}The model M3a4, a) Time-latitude butterfly diagram\nfor the toroidal field in the upper part of the convection zone (color\nimage) and the surface radial magnetic field shown by contours ($\\pm$100G);\nb) the surface variations of the azimuthal velocity (color image)\nand the meridional velocity (contours in the range of $\\pm3.5$ m\/s). }\n\\end{figure}\n\nFigure \\ref{fig:M3a4dr} shows snapshots of the global flows distributions\nin the solar convection zone. In drastic difference to the model M3,\nthe counter-clockwise meridional circulation cell in the upper part\nof the convection zone is divided into two parts. Also, the stagnation\npoint of the bottom cell is shifted equatorward.\n\n\\begin{figure}\n\\includegraphics[width=0.7\\columnwidth]{m8dr}\n\n\\caption{\\label{fig:M3a4dr}The model M3a4: a) the snapshot of angular velocity\nprofile, $\\Omega\\left(r,\\theta\\right)\/2\\pi$, contours are in range\nof 347-450 nHz; b) the streamlines of the meridional circulation;\nc) the radial profile of the meridional flow at $\\theta=45^{\\circ}$;\nd) the profiles of the meridional flow at the specific depths of the\nsolar convection zone.}\n\\end{figure}\n\nFigure \\ref{fig:M3a4} shows the magnetic cycle variations of the\nglobal flows in the Northern segment of the solar convection zone\nfor the model M3a4. The patterns of these variations are qualitatively\nthe same as in the model M3. The model M3a4 show the strong magnetic\ncycle variations of the $\\alpha$ effect. Similar to the model M3\nwe see that the magnetic helicity conservation results into increasing\nthe $\\alpha$ -effect in the subsurface shear layer. It occurs just\nahead of the dynamo wave drifting toward the top. In the polar regions,\nthe $\\alpha$ -effect inverses the sign during inversion of the polar\nmagnetic field. This is different from the model M3 which has the smaller\nstrength of the polar magnetic field than the model M3a4.\n\n\\begin{figure}\n\\includegraphics[width=0.95\\columnwidth]{mf6}\\caption{\\label{fig:M3a4}The same as Figure \\ref{fig:M3} for the model M3a4.}\n\\end{figure}\n\n\\begin{figure}\n\\includegraphics[width=0.8\\columnwidth]{f7a}\n\n\\caption{\\label{fig:avar}a) Variations of the equatorial symmetry (parity\nindex, see, Eq(\\ref{eq:parity})); b) the same as (a) for the mean\ndensity of the toroidal magnetic field flux in the subsurface shear\nlayer. The time series were smoothed to filter out the basic magnetic\ncycle.}\n\\end{figure}\n\n\n\\subsection{The long-term dynamo evolution}\n\nFigure \\ref{fig:avar} shows the smoothed time series of evolution\nof the global properties of the dynamo model, such as the equatorial\nsymmetry index, or the parity index $P$, (see, Eq(\\ref{eq:parity}))\nand the mean density of the toroidal magnetic field flux, $\\overline{B^{T}}$,\nin the subsurface shear layer, see, Eq(\\ref{eq:bt}). In each time\nseries, the basic magnetic cycle was filtered out. The set of models\nshown in Figure (\\ref{fig:avar}) illustrates the effect of variations\nof magnitude of the eddy-diffusivity of the magnetic helicity density,\n${\\displaystyle \\frac{\\eta_{\\chi}}{\\eta_{T}}}$. From results of \\citet{mitra10},\nit is expected that ${\\displaystyle \\frac{\\eta_{\\chi}}{\\eta_{T}}}<1$.\nIn our set of models it is $0.01<{\\displaystyle \\frac{\\eta_{\\chi}}{\\eta_{T}}}<0.3$.\nThe magnetic helicity diffusion affects the magnetic helicity exchange\nbetween hemispheres \\citep{2000JGR...10510481B,mitra10}. Therefore\nit affects an interaction of the dynamo waves through the solar equator.\nIt is found that the increasing of ${\\displaystyle \\frac{\\eta_{\\chi}}{\\eta_{T}}}$\nresults into change of the parity index $P$. The model M4 show the\nsymmetric about equator magnetic field. The magnitude of $\\overline{B^{T}}$\nin the model M4 is larger than in the models M3 and M5. Both models\nM3 and M5 operate in a weak nonlinear regime with $0.13<\\overline{\\beta}<0.24$\n(see, Eq(\\ref{eq:bet}) and Table(\\ref{tab:C})). The model M4 has\na slightly higher $\\overline{\\beta}$. This means that the magnetic\nhelicity diffusion affects the strength of the dynamo. This is in\nagreement with \\citet{guero10}. The smoothed time series of $\\overline{B^{T}}$\nin the model M4 show oscillations at the end of the evolution. This\nis because the period of the symmetric dynamo mode ( $P=1$) is about\n12 years that is larger than the period of the basic magnetic cycle\nfor antisymmetric mode $\\left(P=-1\\right)$ . The latter is about\n10.5 years.\n\nIt is interesting to compare the kinematic and nonkinematic dynamo\nmodels, which are the models M2 and M3. The nonkinematic model M3\nhas the smaller parameters $\\overline{B^{T}}$and $\\overline{\\beta}$\nthan the model M2. Also, it is found that in the kinematic model M2\nthe stationary phase of evolution is the mix of the dipole-like and\nquadrupole-like parity, with the mean $P\\approx-0.9$. In the model\nM3 the mean $P\\approx-1$. Therefore the nonkinematic dynamo regimes\naffect the equatorial symmetry of the dynamo solution \\citep{bran89,p99}.\n{Figure \\ref{fig:avar}b shows another interesting difference\nbetween the nonlinear kinematic and nonkinematic runs. The kinematic\nmodels M2, M4 and M5 show a very slow evolution toward the stationary\nstage. The given effect was reported earlier by \\citet{pip2013ApJ}.\nThe high $R_{m}=10^{6}$ and small diffusivity $\\eta_{\\chi}$ result\nto the long time-scale of establishment of the nonlinear balance in\nmagnetic helicity density distributions.}\n\n\\begin{figure}\n\\includegraphics[width=0.8\\columnwidth]{f7b}\n\n\\caption{\\label{fig:lt}The same as Figure (\\ref{fig:avar}) for the models\nwith different $C_{\\alpha}$, see Table(\\ref{tab:C}).}\n\\end{figure}\n\nFigure (\\ref{fig:lt}) shows results for the nonkinematic dynamo models\nin a range the $\\alpha$-effect parameter $C_{\\alpha}$. The increasing\nof the $C_{\\alpha}$ results into increasing the nonlinearity of the\ndynamo model. The parameter $\\overline{\\beta}$ grows from $0.2$\nto $1$ with the increasing of $C_{\\alpha}$by factor 4. The model\nM3a2 shows the long-term periodic variations of the parity index and\nthe magnitude of the toroidal magnetic field $\\overline{B^{T}}$.\nThese long-term cycles are likely due to the parity breaking because\nof the hemispheric magnetic helicity exchange. We made the separate\nrun where the magnetic helicity conservation was ignored and did not\nfind the long-term cycles solution. These cycles are not robust against\nchanges of $C_{\\alpha}.$ For the case $\\eta_{\\chi}=0.1\\eta_{T}$,\nthey exist in the range $1.5C_{\\alpha}^{(cr)}3C_{\\alpha}^{(cr)}$}. This coincides\nwith the formation of the second meridional circulation cell near\nthe equator. Currently, it is not clear if both phenomena are tightly\nrelated or this is an accident. This will be studied further.\n\nWe show the first results about effects of the large-scale magnetic\nactivity on the heat transport and the heat energy flux from the dynamo\nregion. This was previously discussed in papers of \\citet{1992AA...265..328B}\nand \\citet{2000ARep...44..771P}. In the mean-field framework, the\nmajor contributions of the large-scale magnetic field on the heat\nenergy balance inside the convection zone are caused by the magnetic\nquenching of the eddy heat conductivity and the energy expenses (associating\nwith the heat energy loss and gain) on the large-scale dynamo. These\nprocesses are modeled by the mean-field heat transport equations.\nThe magnetic perturbations of the heat flux in the model M3 are an\norder of $10^{-3}$ of the background value. It is an order of magnitude\nless at the surface because of the screening effect and the smaller\nstrength of the large-scale magnetic field in the upper layer of the\nconvection zone. The heat perturbation screening effect is due to\nthe huge heat capacity of the solar convection zone \\citep{stix:02}.\nResults of the model M3a4 illustrate it better than the model M3.\nThe model M3a4 shows the strong toroidal magnetic field in the bulk\nof the convection zone (see Figure \\ref{fig:m34}b). In the upper\nlayer of the convection zone, the strength of the toroidal field exceeds\nthe equipartition level. Besides this, the heat flux perturbations\nare efficiently smoothed out toward the top of the dynamo region.\nAnother interesting feature is that the weakly nonlinear model M3\nshows the increasing mean heat flux at the maximum of the magnetic\ncycle. In the model with the overcritical $\\alpha$ effect, we find\nthe opposite situation. The solar observations show the increasing\nluminosity during the maximum of the solar cycles \\citep{1999GeoRL26.3613W}.\nThe variation of the photometric brightness of solar-type stars tends\nto inverse the sign with the increasing level of the magnetic activity\n\\citep{2016ASPC504.273Y}. From the point of view of our model, this\nmeans that the effect of the magnetic shadow become dominant when\nthe total magnetic activity is increased. This is a preliminary conclusion.\nAlso, the relationship between the magnetic shadow effect in the large-scale\ndynamo and the stellar surface darkening because of starspots is not\nstraightforward.\n\nFinally, our results can be summarized as follows: \n\\begin{enumerate}\n\\item We constructed the nonkinematic solar-type dynamo model with the double-cell\nmeridional circulation. The role of the turbulent pumping in the dynamo\nmodel should be investigated. This requires a better theoretical and\nobservational knowledge of the solar meridional circulation. \n\\item The torsional oscillations are explained as a result of the magnetic\nfeedback on the angular momentum transport by the turbulent stresses,\nthe effect of the Lorentz force and the magneto-thermal perturbations\nof the Taylor-Proudman balance. The increasing level of the magnetic\nactivity results in separation of the upper meridional circulation\ncell for two parts. \n\\item The model shows the decrease of the dynamo period with the increase\nof the magnetic cycle amplitude. The shape of the strong magnetic\ncycle is more asymmetric than the shape of the weak cycles. \n\\item The magnetic helicity density diffusion and the increase turbulent\ngeneration of the large-scale magnetic field results in the increasing\nhemispheric magnetic helicity exchange, the magnetic parity breaking\nand the Grand activity cycles. The Grand activity cycles exists in\nthe intermediate range of the $\\alpha$ - effect parameter, when $1.5C_{\\alpha}^{(cr)}