\chapter{A Quantum Energy Inequality Involving Local Modular Data} \cite{Much2022} \begin{equation*} \innerp{\psi}{\energydensity(f)\psi} \ge - \epsilon - \norm{\smash[b]{\Delta}_{\smash[t]{\sharp}}^{-1/2} \ft{g}_{\lambda}(K_{\raisebox{5pt}{\footnotesize$\sharp$}}) \energydensity(f) \FockVacuum} \end{equation*} \section{Misc} \todo{Put this somewhere else.} A \emph{Lorentz transform} is a linear automorphism of Minkowski spacetime which preserves the Lorentz bilinear form. Lorentz transforms are usually represented by (real) $4 \times 4$ matrices, with respect to the standard basis. The \emph{Lorentz group} $\FullLorentzGroup$. \begin{equation*} \FullPoincareGroup = \RR^4 \ltimes \FullLorentzGroup \end{equation*} The relativistic transformation law for one-particle states is given by \begin{equation*} \parens[\big]{U(a,\Lambda) \psi}(p) = e^{ia \cdot p} \psi(\Lambda^{-1} p), \quad \psi \in \hilb{H}, (a,\Lambda) \in \ProperOrthochronousPoincareGroup. \end{equation*} The mapping $(a,\Lambda) \mapsto U(a,\Lambda)$ is a (irreducible) unitary representation of the proper orthochronous Poincaré group $\ProperOrthochronousPoincareGroup$ on the one-particle Hilbert space. By applying any of the second quantization functors we obtain a representation on the multi-particle state space. \begin{equation*} \parens[\big]{U(a,\Lambda) \psi}{}_n(p_1,\ldots,p_n) = e^{ia \cdot (p_1 + \cdots + p_n)} \psi_n(\Lambda^{-1} p_1, \ldots, \Lambda^{-1} p_n), \end{equation*} Poincaré covariance \begin{equation} \label{equation:poincare-covariance-local-algebras} U(g) \localalg{\spacetimeregion{O}} U(g)^* = \localalg{g\spacetimeregion{O}} \qquad g \in \ProperOrthochronousPoincareGroup \end{equation} \begin{definition}{Von Neumann Algebra of Local Observables}{} \begin{equation*} \localalg{\spacetimeregion{O}} = \braces{b(\varphi(f)) \mid \text{$b$ bounded}, f \in \realschwartz{M}, \supp f \subset \spacetimeregion{O}}'' \end{equation*} \end{definition} \section{Basic Concepts of Modular Theory} \index{modular!theory} If $\hilb{H}$ is a Hilbert space we shall denote the $C^*$-algebra of all bounded linear operators on $\hilb{H}$ by $\BoundedLinearOperators{\hilb{H}}$. \begin{definition}{Cyclic and Separating Vectors}{} Suppose $\hilb{H}$ is a Hilbert space and $\mathcal{A}$ is a $C^*$-subalgebra of $\BoundedLinearOperators{\hilb{H}}$. A vector $\Omega \in \hilb{H}$ is called \begin{itemize} \item \emph{cyclic}\index{cyclic vector} for $\mathcal{A}$ if the vector set $\mathcal{A} \Omega$ is dense in $\hilb{H}$. \item \emph{separating}\index{separating vector} for $\mathcal{A}$ if the map $A \mapsto A \Omega$ from $\mathcal{A}$ into $\hilb{H}$ is injective. \end{itemize} \end{definition} Occasionally, a vector that is both cyclic and separating is called \emph{standard}\index{standard vector}. Recall that the commutant of a set $\mathcal{S} \subset B(\hilb{H})$ of operators is defined as the set of all operators $T \in B(\hilb{H})$ which commute with all operators $S$ in $\mathcal{S}$. We shall denote the commutant of $\mathcal{S}$ by $\mathcal{S}'$.\nomenclature[A]{$\mathcal{A}'$}{commutant of $\mathcal{A}$} \begin{proposition}{}{cyclic-separating} Let $\hilb{H}$ be a Hilbert space and $\mathcal{A}$ be a $C^*$-subalgebra of $\BoundedLinearOperators{\hilb{H}}$. \begin{enumerate} \item \label{item:first} A vector is cyclic for $\mathcal{A}$ if and only if it is separating for $\mathcal{A}'$. \item \label{item:second} If $\mathcal{A}$ is a von Neumann algebra, then a vector is cyclic and separating for $\mathcal{A}$ if and only if it is cyclic and separating for $\mathcal{A}'$. \end{enumerate} \end{proposition} \begin{proof} First, suppose that $\Omega \in \hilb{H}$ is cyclic for $\mathcal{A}$. If $A'$ is an element of $\mathcal{A}'$ with $A' \Omega = 0$, then $A' A \Omega = A A' \Omega = 0$ for all $A \in \mathcal{A}$. This means that $A'$ vanishes on the dense subspace $\mathcal{A} \Omega$, and thus on all of $\hilb{H}$, i.e.\ $A'=0$. This proves that $\Omega$ is separating for $\mathcal{A}'$. Conversely, suppose that $\Omega$ is separating for $\mathcal{A}'$. We have to show that the closed subspace $\overline{\mathcal{A}\Omega}$ is all of $\hilb{H}$. Let $P$ be the orthogonal projection onto $\overline{\mathcal{A}\Omega}$. Clearly, any element $A$ of $\mathcal{A}$ maps $\overline{\mathcal{A}\Omega}$ into itself. Thus, $PAP=AP$, and the same holds for $A^*$, that is, $PA^*P=A^*P$. Taking the adjoint of the second equation, we get $PAP=PA$. Hence, $P \in \mathcal{A}'$. Now, $P \Omega = \Omega = I \Omega$, where $I$ is the identity operator on $\hilb{H}$ which obviously also belongs to $\mathcal{A}'$, and the assumption that $\Omega$ is separating for $\mathcal{A'}$ implies $P=I$. Consequently, $\overline{\mathcal{A} \Omega} = \hilb{H}$. Statement~\ref{item:second} directly follows from~\ref{item:first} and the fact that $\mathcal{A}'' = \mathcal{A}$ by the Double Commutant Theorem. \end{proof} If $\Omega$ is separating for $\mathcal{A}$, then every element of $\mathcal{A}\Omega$ is of the form $A\Omega$ with a unique $A \in \mathcal{A}$. This allows us to define an (anti-linear) operator $S_0$ in $\hilb{H}$ with domain $\mathcal{A}\Omega$ by \begin{equation} \label{equation:definition-s0} \quad S_0 A\Omega \defequal A^*\Omega \qquad A \in \mathcal{A}. \end{equation} The operator $S_0$ is densely defined if and only if $\Omega$ is cyclic for $\mathcal{A}$. Since the $*$-operation on $\mathcal{A}$ is involutive, the range of $S_0$ coincides with its domain. \begin{lemma}{}{} If $\Omega$ is a cyclic and separating vector for a von Neumann algebra $\mathcal{A}$, then the operator $S_0$ defined by~\eqref{equation:definition-s0} is closable. \end{lemma} \begin{proof} By \cref{proposition:cyclic-separating}, $\Omega$ is also cyclic and separating for the commutant $\vNa{A}'$. Hence we may, in analogy to $S_0$, define another anti-linear operator $F_0$ in $\hilb{H}$ with dense domain $\mathcal{A}' \Omega$ by \begin{equation*} \quad F_0 B\Omega \defequal B^*\Omega \qquad B \in \mathcal{A'}. \end{equation*} By definitions of $S_0$ and $F_0$, we have for every $A \in \mathcal{A}$ and $B \in \mathcal{A}'$ \begin{equation*} \innerp{S_0 A \Omega}{B \Omega} = \innerp{\Omega}{AB \Omega} = \innerp{\Omega}{BA \Omega} = \innerp{F_0 B\Omega}{A \Omega}. \end{equation*} This adjoint identity establishes that $S_0 \subset F_0^*$. (The \enquote{twisted} appearance of the identity is correct, since it involves anti-linear operators on both sides.) The Hilbert adjoint $F_0^*$ of $F_0$ is closed. Hence, we have shown that $S_0$ has a closed extension, and this implies that $S_0$ is closable. \end{proof} \begin{definition}{Tomita operator}{} Suppose $\Omega$ is a cyclic and separating vector for a von Neumann algebra $\mathcal{A}$. The closure $S = \operatorclosure{S_0}$ of the operator $S_0$ defined on $\mathcal{A}\Omega$ by $S_0 A\Omega = A^*\Omega$ for $A \in \mathcal{A}$ is called the \emph{Tomita operator}\index{Tomita operator}\index{operator!Tomita}\nomenclature[S]{$S$}{Tomita operator} for the pair $(\mathcal{A},\Omega)$. \end{definition} It is a well-known fact that closed operators can be decomposed in a similar fashion to the polar coordinate representation $z = e^{i\arg z} \abs{z}$ of a complex number. We state the theorem in its somewhat uncommon variant for anti-linear operators, as this will be our only use case. \begin{theorem}{Polar Decomposition for Anti-Linear Closed Operators}{polar-decomposition} \index{polar decomposition} Let $T$ be an arbitrary closed anti-linear operator in a Hilbert space $\hilb{H}$. Then there exist a positive selfadjoint linear operator $\abs{T}$ and a partial anti-linear isometry $U$ such that \begin{equation*} T = U \abs{T} \qquad \bracks[\big]{\text{in particular, $\Domain{T} = \Domain{\abs{T}}$}}. \end{equation*} The operators $U$ and $\abs{T}$ are uniquely determined given the additional conditions \begin{equation*} \ker\abs{T} = \ker T \qquad (\ker U)^\perp = (\ker T)^\perp \qquad \ran U = \overline{\ran T}. \end{equation*} \end{theorem} Proofs of this statement are contained in~\cite{ReedSimon1} and~\cite{Schmüdgen2012}. When we speak of \emph{the} polar composition of an operator we tacitly assume that the additional conditions ensuring uniqueness are satisfied. Now we are able to introduce the fundamental objects of modular theory. \begin{definition}{Modular Conjugation, Modular Operator}{} Suppose $\vNa{M}$ is a von Neumann algebra acting on a Hilbert space $\hilb{H}$, and suppose $\Omega \in \hilb{H}$ is a cyclic and separating vector for $\vNa{M}$. Let $S$ be the Tomita operator for $(\vNa{M},\Omega)$ and let \begin{equation*} S = J \Delta^{1/2} \end{equation*} be its polar decomposition. The anti-unitary operator $J$ is called \emph{modular conjugation}\index{modular!conjugation}\nomenclature[J]{$J$}{modular conjugation}. The positive selfadjoint operator $\Delta$ is called \emph{modular operator}\index{modular!operator}\index{operator!modular}\nomenclature{$\Delta$}{modular operator}. The pair $(J,\Delta)$ is said to be the \emph{modular data}\index{modular!data}\index{modular!objects} associated to the pair $(\vNa{M},\Omega)$. \end{definition} \todo{clarify why $J$ is anti-unitary} \begin{definition}{Modular Group}{} Adopt the notation of the foregoing definition. The mapping $\RR \ni t \mapsto \Delta^{it}$ is called the \emph{modular group}\index{modular!group} associated to $(\vNa{M},\Omega)$. \end{definition} The modular group is a strongly continuous one-parameter unitary group on $\hilb{H}$. \begin{proposition}{}{modular-data-unitary} Suppose $\vNa{M}$ is a von Neumann algebra acting on a Hilbert space $\hilb{H}$. Let $U$ be a unitary operator on $\hilb{H}$. Then $U\vNa{M}U^*$ is a von Neumann algebra on $\hilb{H}$. Suppose further that $\Omega \in \hilb{H}$ is a cyclic and separating vector for $\vNa{M}$. Then $U \Omega$ is cyclic and separating for $U\vNa{M}U^*$. Let $(J,\Delta)$ be the modular data associated to $(\vNa{M},\Omega)$. Then $(UJU^*,U{\Delta}U^*)$ is the modular data associated to $(U\vNa{M}U^*,U\Omega)$. \end{proposition} \begin{proof} To prove the first assertion, consider any $A \in (U\vNa{M}U^*)''$. By the Double Commutant Theorem~\cite[Theorem 18.6]{Zhu1993}, it suffices to show that $A \in U\vNa{M}U^*$. As $\vNa{M}$ is a von Neumann algebra, this is equivalent to $U^*\! AU \in \vNa{M}''$, again by the Double Commutant Theorem. Let $B \in \vNa{M}'$. It is easy to check that $UBU^* \in (U\vNa{M}U^*)'$. By assumption, $A$ lies in the commutant of $(U\vNa{M}U^*)'$. Thus we find that $[U^*\! AU,B] = U^* [A,UBU^*] U = 0$, as desired. The set of vectors $U\vNa{M}U^* U\Omega = U\vNa{M}\Omega$ is dense in $\hilb{H}$, since it is the image of $\vNa{M} \Omega$ under the homeomorphism $U$. Thus, the vector $U\Omega$ is cyclic for $U\vNa{M}U^*$. Let us show that it is also separating. Suppose $A$ is in $\vNa{M}$ and $UAU^*U\Omega = UA\Omega = 0$ Since unitaries are injective, $A\Omega = 0$. Now $A=0$ follows from the assumption that $\Omega$ is separating for $\vNa{M}$. We have shown that the mapping $UAU^*U\Omega = UA\Omega$ from $U\vNa{M}U^* \to \hilb{H}$ is injective. Let $S = \operatorclosure{S_0}$ be the Tomita operator associated to $(\vNa{M},\Omega)$, and let $S' = \operatorclosure{S'_0}$ be the Tomita operator associated to $(U\vNa{M}U^*,U\Omega)$. Then we have \begin{equation*} (S'_0 U) A \Omega = S'_0 (U A U^*) U \Omega = (U A^* U^*) U \Omega = U A^* \Omega = (U S_0) A \Omega \end{equation*} for all $A \in \vNa{M}$. Consequently, $S'_0 = U S_0 U^*$ as operators with domain $U\vNa{M}\Omega$. Taking the closure, we obtain $S' = U S U^*$. We can write this as $S' = UJU^* U\Delta^{1/2} U^*$, where $S = J \Delta^{1/2}$ is the polar decomposition of the Tomita operator. It is straightforward to check that $UJU^*$ is anti-unitary and $U\Delta^{1/2} U^*$ is positive selfadjoint, and satisfy the additional condition of \cref{theorem:polar-decomposition}. The uniqueness of the polar decomposition implies that $UJU^*$ is the modular conjugation and $U\Delta U^*$ is the modular conjugation associated to the pair $(U\vNa{M}U^*,U\Omega)$. \end{proof} Finally, let us outline how modular theory enters into algebraic quantum field theory. \begin{theorem}{Reeh--Schlieder Theorem}{reeh-schlieder} Let $\spacetimeregion{O}$ be any open spacetime region. Then the vacuum vector $\Omega$ is cyclic for $\localalg{\spacetimeregion{O}}$. If $\spacetimeregion{O}'$ is non-empty, then $\Omega$ is also separating for $\localalg{\spacetimeregion{O}}$. \end{theorem} By Reeh-Schlieder (\cref{theorem:reeh-schlieder}), the vacuum $\Omega$ is cyclic and separating for $\localalg{\spacetimeregion{O}}$. Thus, modular theory \section{The Geometric Action of the Modular Operator Associated With a Wedge Domain} \begin{definition}{Right and Left Wedge, General Wedges}{} The \emph{right wedge}\index{wedge!right}\nomenclature[WR]{$\rightwedge$}{right wedge} and \emph{left wedge}\index{wedge!left}\nomenclature[WL]{$\leftwedge$}{left wedge} in Minkowski space $M$ are the open subsets \begin{equation*} \rightwedge \defequal \braces[\big]{x \in M \vcentcolon x^1 > \abs{x^0}} \quad \text{and} \quad \leftwedge \defequal \braces[\big]{x \in M \vcentcolon x^1 < -\abs{x^0}}. \end{equation*} We say that a spacetime region $W \subset M$ is a \emph{wedge}\index{wedge} if there exists an element $g$ of the Poincaré group such that $W = g \rightwedge$. \end{definition} Instead of the right wedge, we could just as well have used the left wedge to define the notion of a general wedge, since they are transformed into each other by space inversion. \begin{lemma}{}{general-wedge-from-right-wedge} If a spacetime region $W$ is a wedge, then there exists an element $g$ of the proper orthochronous Poincaré group such that $W = g \rightwedge$. \end{lemma} \begin{proof} \todo{xxx} \end{proof} In the standard representation of the Lorentz group, the boost (or velocity transformation) along the $x^1$-axis with rapidity $2 \pi t$ is given by the matrix\footnote{% This matrix depends on the choice of metric signature. Ours is $(+,-,-,-)$. For $(-,+,+,+)$, use \begin{equation*} \Lambda(t) = \begin{pmatrix} \phantom{-}\cosh(2 \pi @ t) & -\sinh(2 \pi @ t) & \; 0 \; & \; 0 \; \\ -\sinh(2 \pi @ t) & \phantom{-}\cosh(2 \pi @ t) & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ \end{pmatrix}. \end{equation*} } \begin{equation} \label{equation:lorentz-boost} \Lambda(t) = \begin{pmatrix} \cosh(2 \pi @ t) & \sinh(2 \pi @ t) & \; 0 \; & \; 0 \; \\ \sinh(2 \pi @ t) & \cosh(2 \pi @ t) & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ \end{pmatrix} \end{equation} The following proposition shows that $t \mapsto \Lambda(t)$ is a one-parameter subgroup of the stabilizer group of the right wedge with respect to the action of the Lorentz group on subsets of Minkowski space. \begin{proposition}{}{} \begin{enumerate} \item $\Lambda(s + t) = \Lambda(s) \Lambda(t)$ for all $s,t \in \RR$. \item $\Lambda(t) \rightwedge = \rightwedge$ for all $t \in \RR$. \end{enumerate} \end{proposition} \begin{proof} The first property can be verified by direct computation. Let us prove the second. By definition, the image $\Lambda(t) x$ of a vector $x \in M$ lies in $\rightwedge$ if and only if \begin{equation*} x^0 \sinh(2 \pi t) + x^1 \cosh(2 \pi t) > \abs{x^0 \cosh(2 \pi t) + x^1 \sinh(2 \pi t)}, \end{equation*} or equivalently \begin{equation*} x^0 \parens[\big]{\sinh(2 \pi t) \mp \cosh(2 \pi t)} + x^1 \parens[\big]{\cosh(2 \pi t) \mp \sinh(2 \pi t)} > 0 \end{equation*} for both sign choices. Using the definitions of the hyperbolic sine and cosine, this may be further simplified to $(x^1 \mp x^0) e^{2 \pi t} > 0$, which holds if and only if $x^1 > \abs{x^0}$, since the exponential is always positive. So we have shown that \begin{equation} \label{equation:image-right-wedge} \Lambda(t) x \in \rightwedge \iff x \in \rightwedge. \end{equation} This implies that $\Lambda(t)\rightwedge \subset \rightwedge$ for all $t \in \RR$. Conversely, given an arbitrary vector $y \in \rightwedge$, we have to find $x \in \rightwedge$ such that $\Lambda(t) x = y$. Consider $x = \Lambda(-t) y$. Clearly, $x \in \rightwedge$, because of $y \in \rightwedge$ and~\eqref{equation:image-right-wedge}. Now it follows from $\Lambda(-t) = \Lambda(t)^{-1}$ that in fact $\Lambda(t) x = y$. \end{proof} It can easily be seen that $\rightwedge' = \leftwedge$, and so \cref{theorem:reeh-schlieder} applies. \begin{theorem}{Bisognano--Wichmann Theorem \textmd{\cite{Bisognano1975}}}{bisognano-wichmann} For the theory of a free scalar field in Minkowski spacetime, let $\spacetimeregion{O} \mapsto \localalg{\spacetimeregion{O}}$ be the net of von Neumann algebras of local observables. If $(J,\Delta)$ is the modular data associated to the algebra $\localalg{\rightwedge}$ of the right wedge and the vacuum $\Omega$, then \begin{equation*} J = \Theta \cdot U\parens[\big]{0, R_{23}(\pi)} \qquad \Delta^{it} = U\parens[\big]{0,\Lambda(t)}, \end{equation*} where $U$ is the theory's unitary representation of the proper orthochronous Poincaré group. \end{theorem} \todo{give definition of $\Theta$ and $R_{23}$} Note that above statement is for the right wedge only. Let us investigate how the modular group changes, if we consider another wedge region. By \cref{lemma:general-wedge-from-right-wedge} any wedge $W$ can be obtained as $W = g\rightwedge$, where $g$ is a proper orthochronous Poincaré transformation. The covariance property~\eqref{equation:poincare-covariance-local-algebras} of $\vNa{R}$ implies \begin{equation*} \localalg{W} = U(g) \localalg{\rightwedge} U(g)^*. \end{equation*} The vacuum $\Omega$ is Poincaré invariant: \begin{equation*} U(g) \Omega = \Omega. \end{equation*} We write $(J_W,\Delta_W)$ for the modular data associated to $(\localalg{W},\Omega)$. By \cref{proposition:modular-data-unitary} \begin{equation*} J_W = U(g) J U(g)^* \qquad \Delta_W = U(g) \Delta U(g)^* \end{equation*} Recall that the modular group $\Delta_W^{it}$ is defined by means of functional calculus. This raises the following problem: given a selfadjoint operator $A$, a unitary operator $U$ and a suitable function $f$ we want to express $f(UAU^*)$ in terms of $f(A)$, if possible. Note that two different functional calculi are at play here, the former is for $UAU^*$ and the latter for $A$. Simple functions such as polynomials suggest $f(UAU^*) = Uf(A)U^*$. That this is generally true is the statement of the following Lemma. \begin{lemma}{}{functional-calclus-unitary-trafo} Suppose that $A$ is a selfadjoint operator on a Hilbert space $\hilb{H}$, with spectral measure $E_A$. Suppose $U$ is an unitary operator on $\hilb{H}$, and let $E_{U\! @AU^*}$ denote the spectral measure of the (selfadjoint) operator $UAU^*$. Then we have $U E_A U^* = E_{U\! @AU^*}$, and \begin{equation*} U f(A) U^* = f(U\! @@AU^*) \end{equation*} for all Borel functions $f : \RR \to \CC$. \end{lemma} \todo{write down the simpler proof} \begin{proof} For each regular value $\lambda \in \rho(A)$ let \begin{equation*} R_A(\lambda) = (A-\lambda)^{-1} \end{equation*} denote the resolvent operator of $A$. This proof is based on Stone's Formula \todo{reference}, which relates the resolvent to the spectral projections of $A$: If $E_A$ is the spectral measure of $A$ and $\alpha < \beta$ are real numbers, then \begin{equation*} \stronglim_{\varepsilon \downarrow 0} \frac{1}{\pi i} \int_{\alpha}^{\beta} \bracks{R_A(\lambda + i \varepsilon) - R_A(\lambda - i \varepsilon)} d\lambda = E_A \parens[\big]{\bracks{\alpha,\beta}} + E_A \parens[\big]{\parens{\alpha,\beta}} \end{equation*} Recall that a spectral measure is countably additive. As a consequence, \begin{equation*} \stronglim_{\alpha \uparrow a} \stronglim_{\beta \downarrow b} \stronglim_{\varepsilon \downarrow 0} \frac{1}{2\pi i} \int_{\alpha}^{\beta} \bracks{R_A(\lambda + i \varepsilon) - R_A(\lambda - i \varepsilon)} d\lambda = E_A \parens[\big]{\bracks{a,b}} \end{equation*} for all $a \in \RR \cup \braces{-\infty}$, $b \in \RR \cup \braces{\infty}$. Observe that $\rho(A) = \rho(U\! @AU^*)$ and that for each (common) regular value $\lambda$ we have \begin{equation*} R_{U\! @AU^*}(\lambda) = U R_A(\lambda) @ U^*\!. \end{equation*} Since conjugation with an unitary commutes with the strong operator limit, we obtain \begin{equation*} E_{U\! @AU^*} \parens[\big]{\bracks{a,b}} = U E_A \parens[\big]{\bracks{a,b}} U^* \end{equation*} for all $a,b \in \RR$. The collection $\mathcal{A}$ of all subsets $S$ of $\RR$ such that $E_{U\! @AU^*} \parens[\big]{S} = U E_A \parens[\big]{S} U^*$ is a $\sigma$-algebra on $\RR$. We have shown that all closed intervals belong to $\mathcal{A}$. It is well known that the Borel-$\sigma$-algebra $\mathcal{B}$ of $\RR$ is generated by the closed intervals. Hence, $\mathcal{B} \subset \mathcal{A}$. This shows that the spectral measures $U E_A U^*$ and $E_{U\! @AU^*}$ coincide. \end{proof} \begin{equation*} \Delta_W^{it} = U(g) \Delta^{it} U(g)^* = U(g) U\parens[\big]{0,\Lambda(t)} U(g)^* = U\parens[\big]{g(0,\Lambda(t))g^{-1}} \end{equation*} Recall that Stone's Theorem \todo{add reference} states that every strongly continuous one-parameter unitary group is of the form $t \mapsto e^{itK}$ with a uniquely determined selfadjoint operator $K$, which is called \emph{infinitesimal generator} of the group. \begin{definition}{Modular Hamiltonian}{} The infinitesimal generator of the modular group associated to a spacetime region $\spacetimeregion{O}$ is called the \emph{modular Hamiltonian}\index{modular!Hamiltonian}\nomenclature[KO]{$K_{\spacetimeregion{O}}$}{modular Hamiltonian for $\spacetimeregion{O}$} for said region, and denoted $K_{\spacetimeregion{O}}$. \end{definition} In other words, $K_{\spacetimeregion{O}}$ is the unique selfadjoint operator such that $\Delta_{\spacetimeregion{O}}^{it} = e^{itK_{\spacetimeregion{O}}}$ for all $t \in \RR$. \begin{proposition}{}{} The modular Hamiltonian for the right wedge is given by $d \Gamma(A)$, where \begin{equation*} A\psi(p) = - \frac{2\pi}{i} \parens[\big]{\partial_0 \psi(p) \, p^1 + \partial_1 \psi(p) \, p^0} \end{equation*} \end{proposition} \todo{domain, proof} \section{Complex Lorentz Transformations} The main result of this section is \cref{proposition:main-result}. By definition, the \emph{complex Lorentz group}\index{Lorentz group!complex}\nomenclature[LC]{$\ComplexLorentzGroup$}{complex Lorentz group} $\ComplexLorentzGroup$ is the isometry group of complex Minkowski space $M+iM \cong \CC^4$ with respect to the inner product \begin{equation*} \innerp{z_1}{z_2} = \innerp{x_1}{x_2} - \innerp{y_1}{y_2} + i \parens[\big]{\innerp{x_1}{y_2} + \innerp{x_2}{y_1}}. \end{equation*} The \emph{complex Poincaré group}\index{Poincaré group!complex}\nomenclature[PC]{$\ComplexPoincareGroup$}{complex Poincaré group} is the semidirect product $\ComplexPoincareGroup \defequal \CC^4 \ltimes \ComplexLorentzGroup$. The action of $\ComplexPoincareGroup$ on $M+iM$ is defined in the obvious way. The complex Poincaré group has just two connected components, the subgroup $\ProperComplexPoincareGroup$ and the subset $\ImproperComplexPoincareTransformations$, differentiated by the sign of $\det \Lambda \in \braces{\pm 1}$ for its elements $(z,\Lambda)$. The (real) proper orthochronous Poincaré group $\ProperOrthochronousPoincareGroup$ is a subgroup of $\ProperComplexPoincareGroup$. Each of the two following sections deals with a subgroup $G$ of $\ProperOrthochronousPoincareGroup$, and the possibility of extending a unitary representation of $G$ to a larger set within $\ProperComplexPoincareGroup$. \subsection{Analytic Continuation of the Space-Time Translation Group} %\todo{a short intro} Let $a \mapsto U(a)$ be a strongly continuous unitary representation of the additive group of $\RR^4$ (on some separable Hilbert space). By a generalization of Stone's Theorem~\cite[Theorem VIII.12]{ReedSimon1}, there exists a unique projection-valued measure $E$ on $\RR^4$ such that \begin{equation} \label{equation:spectral-resolution-translation} U(a) = \int_{\RR^4} \exp(ia \cdot k) \, dE(k) \qquad a \in \RR^4. \end{equation} Then one can define a vector $P$ of unbounded selfadjoint operators \begin{equation*} P_i = \int_{\RR^4} k_i \, dE(k) \qquad i=0,\ldots,3 \end{equation*} which have a common dense domain $D$ and satisfy \begin{equation*} a \cdot P = \int_{\RR^4} a \cdot k \, dE(k) \qquad a \in \RR^4. \end{equation*} We are specifically interested in the representation \begin{equation*} U(a) \defequal U(a,I) \qquad a \in \RR^4 \end{equation*} obtained by restricting the unitary representation of the Poincaré group $\RestrictedPoincareGroup$ on Fock space to the subgroup of spacetime translations. In this case the vector operator $P$ carries the physical meaning of energy-momentum, and we impose the so-called \emph{spectrum condition} \begin{equation*} \langle a \cdot P \rangle_{\psi} \ge 0 \qquad \forall \psi \in D \; \forall a \in \ClosedForwardCone, \end{equation*} where $\ClosedForwardCone \defequal \braces{a \in \RR^4 \vcentcolon a \cdot a \ge 0, a^0 \ge 0}$ is the \emph{closed forward cone}\index{cone!closed forward}\nomenclature[V]{$\ClosedForwardCone$}{closed forward cone}. It can be shown \cite{Uhlmann1961} that the spectrum condition is equivalent to the statement that the support of the spectral measure is contained in the closed forward cone, i.e.\ $\supp(E) \subset \ClosedForwardCone$. Spectral calculus allows us to extend $a \mapsto U(a)$ to complex arguments $z \in \CC^4$ by simply replacing $a$ with $z$ in the spectral resolution~\eqref{equation:spectral-resolution-translation} of $U(a)$. However, one obtains, in general, an unbounded operator. It is a consequence of the spectrum condition that $U(z)$ is bounded whenever $z$ lies in the \emph{closed forward tube}\index{tube!closed}\nomenclature[T]{$\ClosedForwardTube$}{closed forward tube} $\ClosedForwardTube \defequal \RR^4 + i\ClosedForwardCone$. Observe that the set $\ClosedForwardTube$ is closed under vector addition and thus forms a commutative monoid; it is not a group. \begin{proposition}{}{} For every $z \in \ClosedForwardTube$ the operator \nomenclature[U]{$U(z)$}{complex translation} \begin{equation} \label{equation:definition-complex-translation} U(z) \defequal \int_{\ClosedForwardCone} \exp(iz \cdot k) \, dE(k) \end{equation} is bounded. Moreover, $U(w+z) = U(w) U(z)$ for all $w,z \in \ClosedForwardTube$. \end{proposition} \begin{proof} By a general property of spectral integrals~\cite[Proposition 4.18]{Schmüdgen2012}, the operator $U(z)$ is bounded if (and only if) the function $f(k) = \exp(iz \cdot k)$ is bounded $E$-almost everywhere. In view of the fact that $E$ is supported in the closed forward cone $\ClosedForwardCone$, it is sufficient to show that $f$ is bounded on $\ClosedForwardCone$. %the $E$-essential supremum of the function Since $z$ lies in the closed forward tube, $z=x+iy$ with $x \in \RR^4$ and $y \in \ClosedForwardCone$. Now $\abs{f(k)} = \exp(-y \cdot k)$, and on $\ClosedForwardCone$ this is bounded by $1$ because $y \cdot k \ge 0$ for all $k \in \ClosedForwardCone$. The identity $U(w+z) = U(w) U(z)$ follows from $\exp(i(w+z) \cdot k) = \exp(iw \cdot k) \exp(iz \cdot k)$ and the boundedness of the operators, see~\cite[Proposition 4.16(iii) and (v)]{Schmüdgen2012}. \end{proof} \begin{proposition}{}{} If $(b,\Lambda) \in \RestrictedPoincareGroup$ and $z \in \ClosedForwardTube$, then $\Lambda z \in \ClosedForwardTube$ and \begin{equation*} U(b,\Lambda) U(z) U(b,\Lambda)^* = U(\Lambda z). \end{equation*} \end{proposition} \begin{proof} We exploit the uniqueness of the projection-valued measure $E$ satisfying~\eqref{equation:spectral-resolution-translation}. Since $\Lambda$ acts continuously on $\RR^4$, $\Lambda^{-1} S$ is a Borel set whenever $S \subset \RR^4$ is, and \begin{equation*} F(S) \defequal U(b,\Lambda) E(\Lambda^{-1} S) U(b,\Lambda)^* \qquad S \in \BorelSigmaAlgebra{\RR^4} \end{equation*} is a well-defined projection-valued measure on $\RR^4$. By the Transformation Formula for Spectral Integrals~\cite[Proposition 4.24]{Schmüdgen2012}, we have \begin{align} \int_{\RR^4} \exp(iz \cdot k) \, dF(k) &= U(b,\Lambda) \bracks[\bigg]{\,\int_{\RR^4} \!\exp(iz \cdot \Lambda k) \, dE(k)} U(b,\Lambda)^* \nonumber\\ &= U(b,\Lambda) \bracks[\bigg]{\,\int_{\RR^4} \!\exp(i \Lambda^{-1} z \cdot k) \, dE(k)} U(b,\Lambda)^* \nonumber\\ \label{equation:F-integral} &= U(b,\Lambda) U(\Lambda^{-1} z) U(b,\Lambda)^* \end{align} for all $z \in \ClosedForwardTube$. In particular, for $z=a \in \RR^4$ the last term equals \begin{equation*} U(b,\Lambda) U(\Lambda^{-1} a) U(b,\Lambda)^* = U(a) \qquad a \in \RR^4, \end{equation*} which can be seen by applying $U$ to the identity \begin{equation*} (b,\Lambda) (\Lambda^{-1} a, I) (b,\Lambda)^{-1} = (a,I). \end{equation*} Hence, $U(a) = \int \exp(ia \cdot k) dF(k)$ for all $a \in \RR^4$. We conclude $E = F$. Now~\eqref{equation:F-integral} asserts that $U(z) = U(b,\Lambda) U(\Lambda^{-1} z) U(b,\Lambda)$ for all $z \in \ClosedForwardTube$ and a substitution of $z$ by $\Lambda z$ yields the desired identity. \end{proof} %\begin{proposition}{Analyticity of the Complex Translation Monoid}{analyticity-complex-translations} \begin{proposition}{\textmd{\cite[Theorem 4]{Uhlmann1961}}}{analyticity-complex-translations} The operator-valued map $z \mapsto U(z)$ given by~\eqref{equation:definition-complex-translation} is strongly continuous on $\ClosedForwardTube$ and analytic on $\OpenForwardTube$. \end{proposition} \todo{Explain what it means for an operator-valued function of several complex variables to be analytic.} Next we consider an operator-valued tempered distribution $u$ that is \emph{covariant} in the sense that it obeys the relativistic transformation law \begin{equation} \label{equation:covariance-distribution} U(g) u(f) U(g)^* = u(f_g) \qquad g \in \RestrictedPoincareGroup, f \in \schwartz{\RR^4}, \end{equation} where $f_g(x) = f(g^{-1} x)$ for all $x \in M$. In particular, if $g=(a,I)$ is the translation by a vector $a \in \RR^4$, then~\eqref{equation:covariance-distribution} and the invariance of the vacuum vector $\FockVacuum$ imply \begin{equation} \label{equation:real-translation-law} U(a) u(f) \FockVacuum = u(f_a) \FockVacuum \qquad \forall a \in \RR^4. \end{equation} We would like to extend this law to complex translation vectors, but translating a function defined on $\RR^4$ by a complex vector is not a sensible operation. Nevertheless, we have $\FT{f_a}(p) = \exp(ia \cdot p) \ft{f}(p)$ in Fourier space, and $\exp(iz \cdot p) \ft{f}(p)$ is a well defined function of $p \in \RR^4$ even when $z \in \CC^4$. The obvious idea would be to define $f_z$ as the inverse Fourier transform of this function. This does not work because $\exp(iz \cdot p) \ft{f}(p)$ is generally not in the Schwartz class. However, thanks to the spectrum condition we may modify this function outside of the closed forward cone. \begin{lemma}{}{depends-only-on-restriction} Let $u$ be a covariant operator-valued tempered distribution. Then the vector $u(f) \FockVacuum$, where $f \in \schwartz{\RR^4}$, depends only on the restriction of $\ft{f}$ to $\ClosedForwardCone$. \end{lemma} \begin{proof} We consider a Schwartz function $g \in \schwartz{\RR^4}$ and the operator $G = \int g(k) dE(k)$, where $E$ is the unique projection-valued measure on $\RR^4$ such that $U(a) = \int \exp(ik \cdot a) dE(k)$ for all $a \in \RR^4$. Let $g(k) = (2 \pi)^{-2} \int \ift{g}(a) \exp(ia \cdot k) da$ be the Fourier decomposition of $g$. \begin{multline*} \hspace{1cm} (2 \pi)^2 G = \int_{\ClosedForwardCone} \!\int_{\RR} \ift{g}(a) \exp(ia \cdot k) \, da \, dE(k) = \\ = \int_{\RR} \ift{g}(a) \!\int_{\ClosedForwardCone} \exp(ia \cdot k) \, dE(k) \ da = \int_{\RR} \ift{g}(a) U(a) \, da \hspace{1cm} \end{multline*} \question{Darf ich hier wirklich die Integrationsreihenfolge vertauschen?} Recall that the Fourier transform of $u$ is defined by $\ft{u}(f) = u(\ft{f}@@)$ for $f \in \schwartz{\RR^4}$. We obtain the action of the translation group on $\ft{u}(\ft{f}@@)\FockVacuum$ by definition chasing and~\eqref{equation:real-translation-law}: \begin{equation*} U(a) \ft{u}(\ft{f}@@)\FockVacuum = U(a) u(f)\FockVacuum = u(f_a)\FockVacuum = \ft{u}(\ft{f}_a)\FockVacuum = \ft{u}(\ft{f}e_a)\FockVacuum \end{equation*} Here $e_a$ stands for the function $e_a(p) = \exp(ia \cdot p)$. \begin{equation*} G @\ft{u}(\ft{f}@@)\FockVacuum = \int \ift{g}(a) @\ft{u}(\ft{f}e_a)\FockVacuum \, da = \ft{u} \parens[\bigg]{\ft{f} \int \ift{g}(a) e_a da} \FockVacuum = \ft{u}(\ft{f} g) \FockVacuum \end{equation*} The second identity is due to the continuity of the vector-valued map $f \mapsto u(f) \FockVacuum$. If the support of $g$ does not intersect the support of $E$, i.e.\ the closed forward cone, then $G=0$. Thus, $\ft{u}(\ft{f} g) \FockVacuum = 0$. This proves that $u(f_1) \FockVacuum = u(f_2) \FockVacuum$ when $\supp(\ft{f_1} - \ft{f_2}) \subset \ClosedForwardCone$. \end{proof} This fact inspires the definition \begin{equation*} f_z \defequal d_z * f \qquad z \in \ClosedForwardTube \end{equation*} where $d_z$ is any Schwartz function on $\RR^4$ such that $\FT{d_z}(p) = \exp(iz \cdot p)$ for all $p \in \ClosedForwardCone$. Such a function does exist \todo{elaborate, smooth cutoff}. Then $f_z$ will be Schwartz class as well. Moreover, $u(f_z) \FockVacuum$ does not depend on the specific choice of $d_z$, by~\cref{lemma:depends-only-on-restriction}. %\begin{lemma}{}{} %For every $z \in \ClosedForwardTube$ there exists a Schwartz function $d_z \in \schwartz{\RR^4}$ %such that $\ft{e_z} \in \schwartz{\RR^4}$ and $\ft{e_z}(p) = \exp(iz \cdot p)$ for $p \in \ClosedForwardCone$. %\end{lemma} \begin{proposition}{}{} Let $u$ be a covariant operator-valued tempered distribution, and let $f \in \schwartz{\RR^4}$ be a test function. Then we have, in generalization of~\eqref{equation:real-translation-law}, \begin{equation*} U(z) u(f) \FockVacuum = u(f_z) \FockVacuum \qquad \forall z \in T_+. \end{equation*} \end{proposition} \begin{proof} By \cref{proposition:analyticity-complex-translations}, the function $z \mapsto U(z) u(f) \FockVacuum$ is analytic on the open forward tube. \todo{Zeige, dass $z \mapsto u(f_z) \FockVacuum$ ebenfalls analytisch ist. Dann folgt die Behauptung wohl mit Edge of the Wedge~\cite[Theorem 2-17]{Streater1964}} \end{proof} \subsection{Complex Lorentz Boosts} The Lorentz boosts $\Lambda(t)$ given by the matrices~\eqref{equation:lorentz-boost} in standard representation have a natural interpretation for complex parameters, since the hyperbolic functions $\cosh$ and $\sinh$ extend analytically to the whole complex plane. In view of Lemma xxx it follows immediately that the matrix-valued function $\CC \ni w \mapsto \Lambda(w)$ is entire analytic. In particular, the vector-valued function $\CC \ni w \mapsto \Lambda(w) z$ is entire analytic for every fixed vector $z \in \CC^4$. We are particularly interested in the case of a purely imaginary parameter. The relations $\cosh iz = \cos z$ and $\sinh iz = i \sin z$ between the complex hyperbolic and trigonometric functions imply \begin{equation*} \Lambda(is) = \begin{pmatrix} \phantom{i}\cos(2 \pi @ s) & i\sin(2 \pi @ s) & \; 0 \; & \; 0 \; \\ i\sin(2 \pi @ s) & \phantom{i}\cos(2 \pi @ s) & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ \end{pmatrix} \qquad \forall s \in \RR. \end{equation*} \begin{equation*} \Lambda(is) (x+iy) = \begin{pmatrix} \cos(2 \pi @ s) x^0 - \sin(2 \pi @ s) y^1 \\ \cos(2 \pi @ s) x^1 - \sin(2 \pi @ s) y^0 \\ x^2 \\ x^3 \end{pmatrix} +i \begin{pmatrix} \sin(2 \pi @ s) x^1 + \cos(2 \pi @ s) y^0 \\ \sin(2 \pi @ s) x^0 + \cos(2 \pi @ s) y^1 \\ y^2 \\ y^3 \end{pmatrix} \end{equation*} We \begin{equation*} \mathcal{J} \defequal \Lambda(i/2) = \begin{pmatrix} -1 & 0 & \; 0 \; & \; 0 \; \\ 0 & -1 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ \end{pmatrix} \end{equation*} \begin{equation*} \mathcal{J}_{\pm} \defequal \Lambda(\pm i/4) = \begin{pmatrix} 0 & \pm i & \; 0 \; & \; 0 \; \\ \pm i & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ \end{pmatrix} \end{equation*} We now turn to the unitary representation of (real) Lorentz boosts \begin{equation*} V(t) \defequal U \parens[\big]{0,\Lambda(t)} \qquad t \in \RR \end{equation*} on Fock space and aim for an analytic extension similar to the previous section. By Stone's theorem theorem there exists a unique selfadjoint operator $K$ such that \begin{equation*} V(t) = \exp(itK) = \int_{\RR} \exp(it \lambda) \,dE_K(\lambda), \end{equation*} where $E_K$ is the spectral measure on $\RR$ associated to $K$. Now we define \emph{complex Lorentz boosts} to be the operators \nomenclature[V]{$V(z)$}{complex Lorentz boost} \begin{equation*} V(z) \defequal \int_{\RR} \exp(iz \lambda) \,dE_K(\lambda) \qquad z \in \CC. \end{equation*} In contrast to the previous section, we \begin{lemma}{}{} Suppose $A$ is a selfadjoint unbounded operator on some Hilbert space $\hilb{H}$. For each complex number $z$ define the closed normal operator $V(z) = e^{izA}$ by means of functional calculus. Let $g \in \schwartz{\RR}$ be a Schwartz function. \begin{enumerate} \item $V(z) V(w) = V(z + w)$ for all $z,w \in \CC$. \item The operator $g(A)$ is bounded, and its range is contained in the domain of $V(z)$ for all $z \in \CC$. \item The operator $V(z) g(A)$ is bounded for all $z \in \CC$, and has spectral resolution \begin{equation*} V(z) g(A) = \int e^{iz \lambda} g(\lambda) dE_A(\lambda). \end{equation*} \item The function $z \mapsto V(z) g(A)$ is entire analytic. \end{enumerate} \end{lemma} Remember that a \emph{core}\index{operator!core for an} for a closed densely defined unbounded operator $T$ is, by definition, a linear subspace $\mathcal{D}_0$ of its domain $\Domain{T}$ such that the closure of the restriction of $T$ to $\mathcal{D}_0$ coincides with $T$. %symbolically $\overline{T \vert \mathcal{D}_0} = T$. Each core of $T$ is necessarily a dense subspace of $\Domain{T}$, but a dense subspace of $\Domain{T}$ need not be a core for $T$. \begin{lemma}{A Common Core for All Complex Lorentz Boosts}{common-core-for-complex-lorentz-boots} Adopt the notation of the foregoing lemma. The linear subspace \begin{equation*} \mathcal{D}_0 = \Span \braces{\ran g(K) \vcentcolon g \in \schwartz{\RR}} \end{equation*} is a core for $V(z)$ for every $z \in \CC$. \end{lemma} \begin{proof} xxx \end{proof} \subsection{Application to the Energy Density} \bluetext{Achtung: Dieser Abschnitt ist noch roh, lückenhaft und enthält inkonsistente Notation und falsche Aussagen.} The following three Lemmas are variations of the arguments brought forward by~\citeauthor{Bisognano1975} in their proof of \cref{theorem:bisognano-wichmann}. The main difference is that we state xxx and xxx as operator identities without reference to a field operator, and proof xxx for arbitrary Lorentz-covariant operator-valued distributions rather than products of field operators. This generalization is necessary for the application to the energy density. In addition, we provide in \cref{chapter:convolution} a complete proof of the convolution formula for vector-valued distributions. Roughly speaking, the following Lemma asserts that a translation by a complex vector followed by a suitable imaginary boost is again a complex translation. \begin{lemma}{}{} Let $z = x + iy \in \OpenForwardTube$ with $x,y$ real, and suppose $y^3=y^4=0$. \begin{enumerate} \item If $x \in \rightwedge$, then for all $s \in [0,1/4]$ \begin{equation*} \Lambda(is) z \in \OpenForwardTube, \qquad \ran U(z) \subset \dom V(is), \qquad V(is) U(z) = U \parens[\big]{\Lambda(is) z}. \end{equation*} \item If $x \in \leftwedge$, then the above holds for all $s \in [0,-1/4]$. \end{enumerate} \end{lemma} \nomenclature[dom]{$\dom T$}{domain of the operator $T$\nomnorefpage} \nomenclature[ran]{$\ran T$}{range of the operator $T$\nomnorefpage} \begin{proof} xxx \noindent\begin{minipage}{0.5\textwidth} The vector-valued function $\CC \ni s \mapsto \Lambda(is) z$ is entire analytic. In particular it is continuous, and we have shown that it maps the compact subset $[0,1/4]$ into the open set $\OpenForwardTube$. This implies that there exists a connected open neighborhood $N \subset \CC$ of $[0,1/4]$ such that $\Lambda(is) z \in \OpenForwardTube$ for all $s \in N$. \end{minipage} \hfill \begin{minipage}{0.45\textwidth} \begin{center} \begin{tikzpicture}[baseline=10] \draw[->] (-1,0) -- (3,0) node[right] {\footnotesize$\Real s$}; \draw[->] (0,-1) -- (0,2) node[left] {\footnotesize$\Imag s$}; \draw[faunat,thick,{Parenthesis[]}-{Parenthesis[]}] (0,-0.5) -- (0,0.5); \draw[thick,{Bracket[]}-{Bracket[]}] (-0.3pt,0) -- (2,0); \draw (1,1) node {$N$}; \draw (2,0) node[above] {\footnotesize$\tfrac{1}{4}$}; \draw plot [smooth cycle] coordinates {(2.5,0) (2,1) (1,0.7) (0.3,1) (-0.5,0.7) (-0.3,-0.7) (1.6,-0.7)}; \end{tikzpicture} \end{center} \end{minipage} Let $\xi \in \hilb{F}$ be arbitrary, and let $\eta$ be in the common dense domain $\mathcal{D}_0$ of the operators $V(is)$ from \cref{lemma:common-core-for-complex-lorentz-boots}. Then the function $f_1(s) = \innerp{V(is)^* \eta}{U(z) \xi}$ is well-defined, and entire analytic by Lemma xxx. The function $f_2(s) = \innerp{\eta}{U(\Lambda(is) z) \xi}$ is analytic on $N$, by \cref{proposition:analyticity-complex-translations}. By Lemma xxx, $f_1$ and $f_2$ agree in an open real neighborhood $is$. Since $N$ is an open neighborhood of $0$, there is an $\epsilon >0$ such that $i(-\epsilon,\epsilon) \subset N$. It follows that $f_1 \equiv f_2$ on $N$. core \ldots \end{proof} \begin{lemma}{}{} Let $x \in \rightwedge$ \begin{equation*} \stronglim_{\varepsilon \downarrow 0} V(i/4) U(x+i \varepsilon e_0) = U \parens[\big]{V(i/4)x} = \stronglim_{\varepsilon \downarrow 0} V(-i/4) U(\mathcal{J}x+i \varepsilon e_0) \end{equation*} \end{lemma} \begin{proof} xxx \end{proof} \begin{lemma}{}{} Suppose that $u$ is a covariant operator-valued tempered distribution. Let $f \in \schwartz{M}$ with $\supp f \subset \rightwedge$, and let $g \in \schwartz{M}$ be arbitrary. Then \begin{equation*} V(i/2) g(K) u(f) \FockVacuum = g(K) u(f_{\mathcal{J}}) \FockVacuum \end{equation*} \end{lemma} Here, $K$ is the infinitesimal generator of the group $t \mapsto V(t)$ of real Lorentz boosts, $\FockVacuum$ is the Fock vacuum, and $\mathcal{J}$ is the Lorentz transformation given by the diagonal matrix $\diag(-1,-1,1,1)$. \begin{proof} xxx \end{proof} \begin{equation*} \Delta^{-1/2} g(K) \energydensity(f) \FockVacuum = g(K) \energydensity(f^J) \FockVacuum \end{equation*} Die Anwendung auf die Energiedichte $\energydensity$: \begin{proposition}{}{main-result} Suppose $W \subset M$ is any wedge domain, with associated modular operator $\Delta_W$ and modular Hamiltonian $K_W$. Let $f \in \schwartz{M}$ with $\supp f \subset W$, and let $h \in \schwartz{M}$ be arbitrary. Then \begin{equation*} \norm{\Delta_W^{-1/2} h(K_W) \energydensity(f) \FockVacuum} = \norm{h(K) \energydensity(f_{\mathcal{J}g}) \FockVacuum}, \end{equation*} where $K$ is the modular Hamiltonian of the right wedge $\rightwedge$, and $g$ is any element of $\RestrictedPoincareGroup$ such that $W = g \rightwedge$, and $\mathcal{J} = \diag(-1,-1,1,1)$. \end{proposition} In der Ungleichung aus~\cite{Much2022} ist $h$ eine Gauß-Funktion. \section{Calculating Gaussians of the Modular Hamiltonian} coming soon\ldots \chapterbib \cleardoublepage % vim: syntax=mytex